Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 May 18.
Published in final edited form as: Annu Rev Microbiol. 2020 Sep 8;74:431–454. doi: 10.1146/annurev-micro-020518-115546

Molecular Mechanisms of Drug Resistance in Plasmodium falciparum Malaria

Kathryn J Wicht 1, Sachel Mok 1, David A Fidock 1,2
PMCID: PMC8130186  NIHMSID: NIHMS1702326  PMID: 32905757

Abstract

Understanding and controlling the spread of antimalarial resistance, particularly to artemisinin and its partner drugs, is a top priority. Plasmodium falciparum parasites resistant to chloroquine, amodiaquine, or piperaquine harbor mutations in the P. falciparum chloroquine resistance transporter (PfCRT), a transporter resident on the digestive vacuole membrane that in its variant forms can transport these weak-base 4-aminoquinoline drugs out of this acidic organelle, thus preventing these drugs from binding heme and inhibiting its detoxification. The structure of PfCRT, solved by cryogenic electron microscopy, shows mutations surrounding an electronegative central drug-binding cavity where they presumably interact with drugs and natural substrates to control transport. P. falciparum susceptibility to heme-binding antimalarials is also modulated by overexpression or mutations in the digestive vacuole membrane–bound ABC transporter PfMDR1 (P. falciparum multidrug resistance 1 transporter). Artemisinin resistance is primarily mediated by mutations in P. falciparum Kelch13 protein (K13), a protein involved in multiple intracellular processes including endocytosis of hemoglobin, which is required for parasite growth and artemisinin activation. Combating drug-resistant malaria urgently requires the development of new antimalarial drugs with novel modes of action.

Keywords: antimalarial drug resistance, artemisinin-based combination therapy, piperaquine, transport, hemoglobin, endocytosis, pfcrt, k13

1. INTRODUCTION

The persistence of the Plasmodium falciparum parasite, despite decades of clinical research and treatment protocols, has led to a global malaria disease burden that included over 200 million cases in 2018, according to the World Health Organization (157). These cases resulted in an estimated 405,000 deaths, including 272,000 (67%) children under 5 years, with an overwhelming 94% of mortalities occurring in sub-Saharan Africa. As the utilization and effectiveness of experimental vaccines have been very limited, there is a significant reliance on antimalarial drugs for prophylaxis as well as for treatment of infected patients (30). For thousands of years, malaria was treated with natural products found in bark, roots, or leaves of plants. Their active ingredients, however, were identified and used as isolated drug compounds only in the last century. Perhaps the most successful of these medicines was quinine, a quinoline-containing alkaloid from the bark of cinchona trees (101). Developments in the chemical synthesis of drug analogs led to the 4-aminoquinoline quinacrine and, after further toxicological and pharmacological studies, chloroquine. By the 1950s this fast-acting antimalarial was the frontline treatment for malaria. The use of chloroquine as a single-agent therapy continued globally for decades thanks to its efficacy, availability, low toxicity, and affordability (150). Antimalarials are some of the most commonly used medications in tropical regions, where the substantial need for treatment, exacerbated to a degree by incomplete patient compliance, puts immense drug selection pressure on P. falciparum parasites to evolve resistance mechanisms. When chloroquine resistance emerged in the early 1960s, malaria resurged, continuing for decades in most countries (151). High-throughput screening and drug discovery and development work by the US Army led to the discovery of pyrimethamine and the chloroquine analog amodiaquine, as well as the arylaminoalcohols mefloquine and halofantrine. They too suffered declining efficacy from the late 1980s, owing to the spread of resistant parasites. Another chloroquine derivative, piperaquine, developed under the Chinese National Malaria Elimination Program as a bis 4-aminoquinoline that can overcome chloroquine resistance, has more recently been used in combination therapies in Southeast Asia (14).

The use of two or more compounds with different modes of action for malaria treatment is recommended by the World Health Organization, both to provide necessary cure rates and to delay the onset of resistance. Nonetheless, regions with low mixed-strain transmission rates, specifically in Southeast Asia, have historically been the first to show resistance to frontline drugs (86). Indeed, resistance to chloroquine, mefloquine, and sulfadoxine-pyrimethamine initially arose in that region (14, 112). More recently, artemisinin-based combination therapies (ACTs) have been successful in controlling malaria and have saved countless lives, with the global burden showing a 37% reduction from 2000 to 2015 (43). Artemisinin, originally extracted from the Chinese sweet wormwood Artemisia annua, and its derivatives artemether, artesunate, and dihydroartemisinin (DHA) are fast-acting compounds that contain a unique endoperoxide bridge. These are typically combined with a partner drug having a longer half-life such as lumefantrine, piperaquine, mefloquine, amodiaquine, or more recently pyronaridine, as recommended by the World Health Organization for the treatment of uncomplicated P. falciparum malaria. As early as 2008, emerging artemisinin resistance was detected in Southeast Asia, particularly in the Greater Mekong Subregion (GMS) (86, 153). The most problematic situation currently is the very rapid increase in failure rates for DHA + piperaquine, which has been the first-line treatment and the preferred ACT in most of Southeast Asia (106, 125, 142). The rise of these new DHA-and-piperaquine-resistant parasites threatens the recent progress made in malaria reduction and highlights the need for new interventions.

One initial consideration was to use the prophylactic combination drug atovaquone-proguanil (83). One caveat is that resistance to atovaquone can be readily acquired via mutations in cytochrome b, although these mutant parasites might no longer be transmissible (48). Proguanil is also of limited potency, and its active metabolite cycloguanil can encounter resistance via point mutations in the target dihydrofolate reductase (123). Mechanisms underlying resistance to atovaquone-proguanil and the antifolate combination pyrimethamine-sulfadoxine have recently been reviewed separately (53, 127). Our article instead focuses on modes of action of and mechanisms of resistance to ACT drugs.

2. MOLECULAR MECHANISMS OF THE 4-AMINOQUINOLINES AND ARYLAMINOALCOHOLS

2.1. Mode of Action of the 4-Aminoquinolines and Arylaminoalcohols

Since chloroquine resistance first emerged independently in Southeast Asia and South America, later spreading from Asia to Africa, scientists have been working to understand the mechanistic basis of drug action and resistance (115) (Supplemental Table 1). This work has converged on degradation of host hemoglobin and the subsequent detoxification of heme products. During the asexual blood stage of its life cycle, the developing trophozoite ingests up to 75% of the available hemoglobin, a major cytosolic host erythrocyte protein, by a process of endocytosis via cytostomes. Vesicles containing hemoglobin are transported to the digestive vacuole (DV), an acidic secondary lysosome with a pH of ~5.2 (70). In the DV, the parasite must degrade hemoglobin to acquire the amino acids required for its growth and maturation, a catabolic process mediated by multiple proteases (47). These include the aspartic proteases plasmepsins 1, 2, 4, and 3 (the latter is also known as histo-aspartic protease); the falcipain cysteine proteases; and the zinc protease falcilysin. This process produces denatured globin and a heme by-product, iron protoporphyrin IX [Fe(II)PPIX]. Fe(II)PPIX can be auto-oxidized by O2 to cytotoxic Fe(III)PPIX, which is capable of lipid peroxidation. To mitigate this toxicity, Fe(III)PPIX is biomineralized into hemozoin, an inert and highly insoluble crystalline material (Figure 1a). This brown birefringent crystal is formed mostly during the trophozoite stage, before merozoites develop (132).

Figure 1.

Figure 1

Mechanism of action of quinolines and artemisinins and resistance mechanisms mediated by PfCRT, PfMDR1, and K13 in P. falciparum asexual blood stage parasites. (a) During its asexual blood stage, the P. falciparum parasite, surrounded by its PVM, develops within the host red blood cell. (❶) Quinoline-based antimalarials, including chloroquine, amodiaquine, and piperaquine, concentrate from the parasite cytosol (neutral pH of ~7) into the DV (acidic pH of ~5.2). (❷) Once inside the DV, these weak-base drugs are protonated, existing mostly as pH-trapped charged species that are unable to passively diffuse out through the DV membrane. (❸) As additional molecules diffuse into the DV, their protonated forms bind to the high concentrations of toxic free heme by-product that result from the degradation of host hemoglobin, as well as to grooves on the surfaces of growing hemozoin crystals. The combination of pH trapping and heme binding accounts for the >1,000-fold drug accumulation inside the DV. (❹) The DV membrane protein PfCRT is believed to be involved in transporting peptides released from hemoglobin digestion into the parasite cytosol. (❺) In drug-resistant parasites, mutations in PfCRT enable the efflux of protonated drug molecules out of the DV, away from their heme target. (❻) Mutations in the DV membrane transporter PfMDR1 can also influence parasite susceptibility to these compounds and are thought to enable transport of drugs such as halofantrine, lumefantrine, and mefloquine into the DV, away from their primary site of action. (b) (❶) Artemisinin drugs are activated by cleavage of their endoperoxide by iron protoporphyrin IX (Fe2+-heme), a product of parasite-digested hemoglobin. (❷) The Fe2+-heme-artemisinin carbon-centered radicals alkylate and damage a multitude of parasite proteins, heme, and lipids and inhibit proteasome-mediated protein degradation. K13 mutations, located primarily in the β-propeller kelch domain, confer artemisinin resistance in young rings. (❸) The loss of K13 function provided by mutations has been shown to cause reduced endocytosis of host hemoglobin and (❹) to extend the duration of ring-stage development, perhaps via PK4-mediated eIF2α phosphorylation. These changes result in lowered levels of hemoglobin catabolism and availability of Fe2+-heme as the drug activator and lead to reduced activation of artemisinin drugs. (❺) K13 mutations may activate the unfolded protein response, maintain proteasome-mediated degradation of polyubiquitinated proteins in the presence of artemisinins, and (❻) remove drugs and damaged proteins through an increase in PI3K-mediated vesicular trafficking. (❼) K13 may also help regulate mitochondrial physiology and maintain membrane potential during drug-induced ring-stage quiescence. The asterisk signifies the activated form of ART. Abbreviations: ADQ, amodiaquine; ART, artemisinin; CQ, chloroquine; DV, digestive vacuole; eIF2α, eukaryotic initiation factor 2α; Hb, hemoglobin; HF, halofantrine; LMF, lumefantrine; MFQ, mefloquine; PfCRT, P. falciparum chloroquine resistance transporter; PfMDR1, P. falciparum multidrug resistance 1 transporter; PPQ, piperaquine; PVM, parasitophorous vacuole membrane; RBC, red blood cell.

Multiple studies have established that chloroquine and its 4-aminoquinoline analogs inhibit hemozoin formation in the DV, causing DV swelling and pigment clumping. Their activity can be reversed by inhibiting hemoglobin proteases (49, 131). 4-Aminoquinolines have also been shown to inhibit the formation of β-hematin (synthetic hemozoin) in extracellular biomimetic assays (96). Furthermore, parasitized red blood cells treated with chloroquine, amodiaquine, or piperaquine show a dose-dependent increase in intracellular free heme with a corresponding decrease in hemozoin (26, 32). Chloroquine inhibits hemozoin formation by binding to the fastest-growing crystal face, possibly as a chloroquine-hematin complex (49, 100). The ability of quinoline antimalarials to strongly bind to Fe(II)PPIX, combined with pH trapping, allows for their >1,000-fold accumulation in the DV (16).

The arylaminoalcohols include antimalarials such as quinine, mefloquine, and lumefantrine. Generally, their mechanism of action is not as well understood as that of the 4-aminoquinolines; however, they may also partially interfere with hemozoin formation and detoxification of hemoglobin degradation by-products as a secondary aspect of their modes of action. Indeed, the arylaminoalcohols have been shown to form 1:2 drug:hematin μ-oxo dimer complexes in a similar manner to that of the 4-aminoquinolines, albeit with weaker binding constants (37). In a cell fractionation assay, there was a statistically significant increase in free heme and a decrease in hemozoin for parasites treated with quinine, mefloquine, and lumefantrine at 2.5 times their 50% inhibitory concentration (IC50) values. However, these effects were not as pronounced as the divergent heme/hemozoin levels observed after treatment with chloroquine (27).

2.2. PfCRT as a Primary Driver of Resistance to Chloroquine and Amodiaquine

Chloroquine-resistant parasites display decreased accumulation of chloroquine in the DV, mediated via an energy-dependent drug efflux mechanism (73). The main causal determinant was reported in 2000, when the analysis of a genetic cross between chloroquine-sensitive and chloroquine-resistant strains identified mutations in the P. falciparum chloroquine resistance transporter gene, pfcrt (40). The 49-kDa drug/metabolite transporter protein PfCRT localizes to the DV membrane, consistent with its role in mediating chloroquine efflux out of the DV away from its heme target (29, 40, 81, 115). Very recently, single-particle cryogenic electron microscopy (cryo-EM) determined the PfCRT structure to 3.2 Å resolution, after successful screening of PfCRT-specific antigen-binding fragments (Fabs) that formed complexes with PfCRT as a means to overcome cryo-EM size and analytical limitations (67). The transporter contains 424 amino acids and 10 transmembrane helix domains arranged as five helical pairs that form two-helix hairpins with an inverted antiparallel topology, typical of drug/metabolite transporters, as well as two juxtamembrane helices. Together they form a negatively charged central cavity of around 3,300 Å3 that is open to the DV side in the solved antibody-bound conformation (Figure 2). This cavity is believed to accommodate positively charged drugs and other compounds/solutes that concentrate in the DV, allowing for their transport to the cytosol as PfCRT alternates conformations during its transport cycle. trans-Stimulation efflux experiments have demonstrated that extracellular chloroquine stimulates efflux of preloaded intracellular chloroquine, supporting the transporter or carrier model, as opposed to the channel model, for PfCRT-mediated drug efflux (117). The kinetics, efficiency, and mechanisms of this process are highly dependent on the specific mutations that give rise to a particular PfCRT isoform (16, 19, 67). Earlier studies reported genetic transformation via pfcrt allelic exchange and confirmed that chloroquine-resistant mutations in pfcrt-modified clones are sufficient for producing chloroquine-resistant phenotypes on different genetic backgrounds (122). An important and ubiquitous amino acid substitution in chloroquine-resistant alleles, regardless of origin, is lysine to threonine at position 76. This K76T mutation is always accompanied by multiple additional region-specific mutations. For example, the chloroquine-resistant South American 7G8, African GB4, and Southeast Asian Dd2 PfCRT isoforms harbor five, six, and eight mutations, respectively, compared to the chloroquine-sensitive wild-type 3D7 isoform. The cryo-EM structure of PfCRT was solved for the 7G8 isoform, revealing that all five of its mutations, namely, C72S, K76T, A220S, N326D, and I356L, line the central drug-binding cavity. A minimum of four of these mutations is required to confer chloroquine resistance, suggesting a codependent role for these additional amino acid substitutions (42). Although the 7G8, GB4, and Dd2 PfCRT isoforms are all considered chloroquine resistant, the South American SVMNT PfCRT haplotypes, as opposed to the African/Asian CVIET haplotypes, are equally associated with resistance to amodiaquine (116). Therefore, cross-resistance between chloroquine and amodiaquine is mainly observed on 7G8 backgrounds.

Figure 2.

Figure 2

The structure of 7G8 PfCRT solved by cryogenic–electron microscopy showing the ten transmembrane helices (TM1–10) and two juxtamembrane helices (JM1–2) viewed (a) from the side and (b) from the DV into the central drug-binding cavity. Mutations associated with drug resistance are colored for piperaquine (blue), chloroquine (gold ), or amodiaquine (gold and bold). The substituted amino acids are represented as solid-colored sticks when viewed (b) from the DV only. Abbreviations: ADQ-R, amodiaquine resistance; CQ-R, chloroquine resistance; DV, digestive vacuole; PfCRT, P. falciparum chloroquine resistance transporter; PPQ-R, piperaquine resistance.

Transport studies have revealed that isoforms of PfCRT, expressed in Xenopus oocytes (10, 81), Saccharomyces cerevisiae yeast (8), or proteoliposomes (66), show a dose-dependent chloroquine or quinine uptake only when these resistance-conferring mutations are present, with K76T being a key transport requirement. This proton-coupled transport, which is dependent on a pH gradient and a positive membrane potential, is inhibited by other 4-aminoquinolines, including amodiaquine and piperaquine. Transport is also inhibited by the chloroquine resistance reversal agent verapamil, which presumably competes for the PfCRT drug-binding site (67). PfCRT, which is essential to the parasite, might itself constitute a target of 4-aminoquinolines, which may act as competitive inhibitors of natural PfCRT substrates. Indeed, parasites expressing lower levels of mutant PfCRT are more sensitive to chloroquine, despite showing similar levels of chloroquine accumulation (74). Although the native function of PfCRT is not confirmed, studies have suggested its involvement in mediating the transport of hemoglobin-derived peptides and amino acids out of the DV, as well as glutathione, Cl ions, H+ ions, and iron (7, 73, 75, 81, 103, 162). PfCRT variants expressed in proteoliposomes transport basic amino acids such as arginine, lysine, and histidine, which would be positively charged upon entering PfCRT’s cavity from the acidic DV (66). Notably, lumefantrine and atovaquone or neutral amino acids such as leucine do not compete for the drug-binding site, suggesting PfCRT’s specificity for cationic or protonatable 4-aminoquinoline compounds (67).

Further evidence of PfCRT’s role in transporting essential nutrients from the DV has been reported in metabolomics studies that evaluated the effects of mutations on peptide levels in parasite extracts. Metabolic quantitative trait locus analysis, which links the genome-wide contribution of individual alleles to metabolite concentration, demonstrates a correlation between elevated hemoglobin-derived peptides and the chloroquine-resistance-conferring pfcrt locus (75). Chloroquine-resistant K76T parasites show the largest accumulation of these peptides, such as PEEK, and a significant fitness disadvantage owing to compromised hemoglobin metabolism. Additional PfCRT mutations, including C101F and L272F, which arose in vitro under amantadine and blasticidin selection, respectively, show massively enlarged DVs in trophozoite and schizont stages and confer substantial fitness costs (108). The addition of these mutations on a chloroquine-resistant background causes reversal of chloroquine resistance, presumably by inhibiting PfCRT-mediated chloroquine transport, in agreement with chloroquine uptake data from Xenopus oocytes (32, 72, 108). Furthermore, parasites harboring an L272F mutation confer methionine auxotrophy to chloroquine-resistant Dd2 parasites, which become unable to metabolize and access this necessary amino acid (72).

2.3. PfMDR1 as a Modulator of P. falciparum Susceptibility to 4-Aminoquinolines and Arylaminoalcohols

Another factor that plays a role in resistance to heme-targeting antimalarials is the P-glycoprotein homolog PfMDR1 (also known as Pgh1), encoded by the P. falciparum multidrug resistance 1 transporter gene pfmdr1 (109, 140). Like PfCRT, PfMDR1 lies on the DV membrane, but transport is predicted to be inwardly directed toward the DV (Figure 1a). Topologically, PfMDR1 resembles a typical P-glycoprotein-type ABC transporter, containing two membrane-spanning homologous domains, each consisting of six predicted helices followed by a hydrophilic nucleotide-binding pocket (118). This binding domain appears to be located on the cytosolic side of the DV, where it could first interact with antimalarials. Reducing pfmdr1 copy number increases parasite susceptibility to mefloquine, halofantrine, lumefantrine, quinine, and artemisinin derivatives (121). The amino acid substitution N86Y, common in African strains, modulates drug susceptibility by enhancing parasite resistance to chloroquine and amodiaquine, while increasing susceptibility to mefloquine, lumefantrine, and DHA (144, 145). The use of artesunate + amodiaquine has been found to select for PfMDR1 N86Y and D1246Y in parasites that emerged after therapy, with decreased sensitivity to monodesethyl-amodiaquine, the active metabolite of amodiaquine (28, 98). PfMDR1 mutations in 7G8 South American parasites have more influence on chloroquine resistance than those in Asian Dd2 or African GB4 parasites, suggesting important isoform-specific relationships between PfCRT and PfMDR1.

Optimal access of drugs to their target is essential for their high activity. Mutations in PfMDR1 presumably inhibit the transport of antimalarials from the cytosol to the DV, thereby decreasing the concentration of heme-targeting drugs such as chloroquine and amodiaquine in the DV. On the other hand, antimalarials such as mefloquine, lumefantrine, and halofantrine that are likely to inhibit targets outside the DV become more potent when their transport by PfMDR1 away from the cytosol is restricted (111). Studies with pfmdr1 complementary RNA–injected Xenopus oocytes revealed that PfMDR1 transports chloroquine, quinine, and halofantrine, with single-nucleotide polymorphisms (SNPs) affecting substrate specificity (118). For example, the mutation N86Y results in a transporter, Pgh-1Dd2, that is unable to transport quinine and chloroquine but instead gains a halofantrine-transporting capability. It is possible that some antimalarial drugs, including quinine and mefloquine, not only are Pgh-1 substrates but also inhibit its regular function by occupying a common drug-binding site, thereby blocking the transport of other solutes (111, 118). In Africa, pfmdr1 gene amplification is very rare, probably owing to the less frequent drug pressure and the more frequent mixed infections that reinforce the impact of its known fitness cost, when compared with the situation in Asia, where fitness appears to be less of a dominating factor (113). The N86Y mutant and D1246 wild-type variants have the largest selective advantage in Uganda; however, SNPs in this gene generally have less impact on fitness than does copy number (99).

2.4. Piperaquine Resistance Mechanisms

For several countries, particularly China, Vietnam, and Malawi, that have effectively prevented the use of chloroquine monotherapy, in vitro and in vivo parasite sensitivities to chloroquine have been frequently restored, along with a decrease in certain resistance-associated pfcrt and pfmdr1 mutations (41, 52). Although the quinoline antimalarials are no longer used as monotherapies in most parts of the world, the widely adopted ACTs often rely on them as partner drugs. The earlier use of artesunate + mefloquine as the first-line therapy in some Southeast Asian countries resulted in selection for multicopy pfmdr1, associated with decreased efficacy, and led to the regional adoption of DHA + piperaquine (14, 28, 145). However, the decreased potency of the artemisinin derivatives in the GMS has augmented pressure on the partner drugs. By 2015, piperaquine resistance was being observed in Cambodia (2, 21, 38). In standard dose-response assays, piperaquine-resistant parasites often have biphasic or incomplete growth inhibition curves, complicating the calculation of the IC50 or IC90 values. A robust alternative has been the piperaquine survival assay, in which synchronized ring-stage parasites are exposed to a pharmacologically relevant dose of 200 nM piperaquine for 48 h followed by drug-free culture for an additional 24 h. Survival ratios are measured after 72 h and are calculated as the percentage of the ratio of drug-treated versus mock-treated live parasites (38). Survival rates ≥10% are associated with an increased risk of piperaquine treatment failure and provide a threshold for determining piperaquine resistance in vitro (155).

Defining the genetic basis for piperaquine resistance has been a high priority, motivating genome-wide association studies (GWASs) that have uncovered several potential molecular markers. These include the amplification of the plasmepsin 2 and 3 genes (pm23), which are involved in hemoglobin degradation (3, 15, 93, 155). It was proposed that overproduction of these plasmepsins might interfere with piperaquine’s inhibition of heme detoxification processes in the DV (155). However, it was not possible to confirm whether plasmepsins were directly involved in mediating piperaquine resistance or whether they were compensatory mechanisms for fitness disadvantages of the actual resistance determinant(s) (78). Another study on 78 western Cambodian isolates sampled between 2011 and 2014 strongly suggested that other loci must be responsible for piperaquine resistance (102). Duru et al. (38) isolated artemisinin-resistant parasites that recrudesced from patients treated with DHA + piperaquine and showed piperaquine survival rates ≥10% correlating with single-copy pfmdr1 and several novel mutations in PfCRT, namely H97Y, M343L, or G353V, which arose on the Dd2 allele (Figure 2). Independently, genome-wide SNPs from 183 Cambodian isolates revealed an additional PfCRT mutation, F145I, that was associated with decreased piperaquine sensitivity in isolates harboring amplified pm23 (1).

Evidence that PfCRT mutations might contribute to piperaquine resistance was initially provided when a PfCRT mutation, C350R, from French Guiana field isolates was found to be associated with decreased piperaquine susceptibility. This phenotype was confirmed via genetic editing in the 7G8 line, which resulted in a small but significant ~1.5-fold shift in the piperaquine IC50 value (105). Additionally, a SNP in pfcrt giving rise to the C101F substitution, identified from an in vitro piperaquine selection experiment, was genetically edited in Dd2 parasites and gave an ~140-fold IC90 increase relative to the Dd2 edited isogenic control, a biphasic dose-response curve, and a 3-fold increase in the 50% lethal dose (LD50) (32). Another study, confirmed via pfcrt-modified parasites engineered in Dd2, found that the PfCRT mutations F145I, M343L, and G353V led to ≥10% piperaquine survival ratios and significant IC90 shifts, even without pm23 amplification (114). Intriguingly, these mutations, which all occurred on chloroquine-resistant lines, caused a partial to full reversal of chloroquine resistance, highlighting the major influence of subtle physicochemical changes within the PfCRT transporter on drug associations and phenotypes (32, 105, 114). The prevalence of these different point mutations in the field appears to be driven largely by the relative parasite survival that they confer under a given drug pressure within the region, together with the fitness cost of the particular mutation (51, 142). Two PfCRT mutations that have expanded the most rapidly in the field in the last five years are T93S and I218F, which also line PfCRT’s drug-binding cavity (Figure 2). These point mutations were recently shown via pfcrt-edited parasites to individually give rise to ~10% parasite survival rates under high piperaquine concentrations, while affording reasonable fitness levels relative to the more highly piperaquine-resistant PfCRT mutation F145I (33).

PfCRT’s interactions with piperaquine and chloroquine have been studied by exposing variant isoforms of PfCRT harboring piperaquine-resistance-associated mutations to radiolabeled piperaquine or chloroquine, either in nanodiscs to assess drug binding or in proteoliposomes to measure drug uptake (67). Both chloroquine and piperaquine show binding within the 0.1–0.2 μM Kd range at pH 5.5, regardless of isoform; however, drug uptake is significantly greater in resistant isoforms. For example, the chloroquine-resistant, piperaquine-sensitive 7G8 and Dd2 isoforms show the largest chloroquine uptake and smallest piperaquine uptake, while those with an F145I or C350R mutation show the opposite effect, relative to the wild type, which gives negligible uptake of both drugs. These findings reiterate the specificity of PfCRT for differentially mediating resistance to structurally distinct compounds, and they confirm that PfCRT point mutations give rise to contrasting drug phenotypes. This phenomenon could be leveraged by employing antimalarials such as chloroquine and piperaquine in combination to exert opposing selective pressure on parasites (67). A longer-term approach would be to reverse resistance by identifying specific inhibitors of the PfCRT substrate-binding cavity that prevent it from transporting natural substrates and drugs (124).

3. MECHANISM OF ARTEMISININ ACTION AND RESISTANCE IN P. FALCIPARUM

3.1. Mode of Action of Artemisinins

Since the early 2000s, ACTs have been the first-line treatment for malaria, having quickly been adopted worldwide (151). The four main drug combinations comprising artemisinin derivatives are artesunate + mefloquine and DHA + piperaquine, used in Southeast Asia, and artemether + lumefantrine (Coartem) and artesunate + amodiaquine, used in Africa. In the 1970s, a team led by Dr. Youyou Tu defined artemisinin’s medicinal property and solved its chemical structure, for which she earned a Nobel Prize in Medicine (139). The compound’s remarkable characteristics include rapid drug-activated killing of both the asexual blood stage and early sexual gametocyte forms of P. falciparum parasites within hours of exposure at low nanomolar concentrations, despite having a short half-life of <1 h.

Most antimalarials such as sulfadoxine-pyrimethamine, atovaquone, and chloroquine inhibit either a single target or a single pathway, e.g., DHFR-mediated folate synthesis by sulfadoxine-pyrimethamine, atovaquone inhibition of cytochrome bc1, and heme detoxification by chloroquine. On the other hand, artemisinins have been reported to bind to a very broad array of parasite proteins and appear to affect a multitude of organellar and cellular processes including hemoglobin endocytosis, glycolysis, protein synthesis and degradation, and cell cycle regulation (17, 64, 119, 148). This unique property is due to the cleavage of its endoperoxide bridge by free Fe(II)PPIX liberated from digested hemoglobin. Once activated, the heme-drug carbon-centered radical alkylates heme, proteins, and lipids, which accelerates the generation of more cytotoxic reactive oxygen species via a cluster bomb effect that eventually leads to cell death (64, 148) (Figure 1b). Independent studies have suggested that artemisinins may also target mitochondrial function by depolarizing this organelle’s membrane potential (77, 147).

3.2. Origins, Spread, and Prevalence of Artemisinin Resistance

Delayed parasite clearance following artemisinin treatment was first observed in western Cambodia, the epicenter of emerging antimalarial multidrug resistance (36, 97). Clinical sites that had more than 30% of cases with microscopically positive parasites evident 72 h after initiating artemisinin or ACT treatment were categorized as areas of resistance by the World Health Organization. Subsequently, clinical artemisinin resistance was redefined as parasite clearance half-life (PCT1/2) > 5 h. The PCT is the time required for the parasite density to be reduced by 50% along the log-linear portion of the normalized parasite clearance curve (152). This was the first instance where antimalarial resistance was recognized as a prolonged time to complete parasite clearance rather than an elevated dose of drug required to eliminate the parasite. Resistance has now been documented across multiple countries in the GMS, including Cambodia, Vietnam, Myanmar, Thailand, and Laos (86). This geographic diversification was attributed to artemisinin-resistant parasites disseminating through population migration as well as de novo emergence in new sites, abetted by permissive infections in local species of Anopheles mosquitoes (5, 126, 136). To date, clinical evidence of artemisinin resistance has not been robustly documented outside of Southeast Asia (28).

Resistance to artemisinin also spells a potential disaster for other antimalarial drugs, as it places increased pressure for these partner drugs to work quickly and effectively once artemisinin loses its efficacy. To combat artemisinin resistance, two malaria control strategies have been developed: triple ACTs (TACTs) and mass drug administration (MDA) (34, 146). While TACTs leverage on combining drugs that differ in their modes of action to prevent multidrug resistance from arising or eliminating infections resistant to one of the ACT partner drugs, MDA aims to eliminate pockets of asymptomatic malaria that serve as reservoirs for transmission and persistence of resistant parasites. Clinical testing of TACT efficacy is underway in the second-phase Tracking Resistance to Artemisinin Collaboration II (TRAC II) multiple-site study. Early indications are that the DHA + piperaquine + mefloquine and the artemether + lumefantrine + amodiaquine combinations are promising and may help delay the onset of artemisinin resistance or restore antimalarial sensitivity in areas that were once artemisinin resistant (143).

3.3. k13 as the Primary Determinant of Artemisinin Resistance and Impact on Fitness

In vitro artemisinin resistance is determined by performing a ring-stage survival assay (RSA) where young, 0–3 h postinvasion rings are exposed for 6 h to a pharmacologically relevant 700-nM concentration of the active artemisinin metabolite DHA. In artemisinin resistance, the loss of efficacy in early rings of resistant isolates has been characterized by a survival rate >1% without a change in sensitivity at the mature asexual blood stages (154). The parasite’s k13 gene (kelch13) as a genetic determinant of artemisinin resistance was first identified in a laboratory-based in vitro evolution study. By increasing artemisinin pressure stepwise on the Tanzanian strain F32 over 125 repeated drug cycles spanning 5 years, Ariey et al. (4) identified the M476I mutation in K13 and the D56V mutation in the DNA-directed RNA polymerase II subunit RPB9 as early genetic changes that emerged after ~30 cycles of artemisinin pressure. Besides k13, mutations in six other genes also arose during further in vitro drug selection. This included a stop codon in falcipain 2 and SNPs in a cysteine protease involved in hemoglobin digestion, protein kinase PK7, gamete antigen 27/25 Pfg27, and two genes encoding unknown proteins.

A GWAS by several groups also implicated k13 as a candidate marker of clinical artemisinin resistance (22, 89, 135). In the GWAS of ~1,500 clinical samples from the multiple-site TRAC study, ~26 mutations in the K13 protein were associated with delayed clearance, defined as a PCT1/2 > 5 h (5). While the k13 gene was the only gene that converged using the two powerful approaches, the in vitro–derived M476I mutation has been observed only at a very low frequency of ~0.3% in one field study (158), providing evidence that in vitro selections can identify the correct gene but that the mutations can differ from those that succeed in the field. The latest extended pooled analyses of 3,250 isolates from a compilation of literature and new data by the WorldWide Antimalarial Resistance Network found 20 β-propeller mutations that were associated with a prolonged PCT1/2 in Asia, but this was not true for S522C, A578S, and Q613L mutations, which were observed in Africa (158). Only one nonpropeller mutation, E252Q, located in the apicomplexan-specific domain has been associated with a 1.5-fold-longer half-life, reaffirming the dominant role of propeller mutations in modulating artemisinin responses (158). Likewise, not all propeller mutations confer artemisinin resistance; for example, A578S was detected at low prevalence in Africa but did not confer resistance in vitro, at least in Dd2 parasites (87). Hence, the presence of propeller mutations cannot be used to predict the resistance phenotype. The reverse genetics approach of using gene-specific zinc-finger nucleases or CRISPR/Cas9 site-directed gene editing has been useful in proving the causal association of the most common K13 propeller mutations (C580Y, R539T, I543T, and Y493H) with in vitro artemisinin resistance (RSA >1%) in Asian parasites (44, 130). The level of resistance differs between K13 mutations even on the same parasite background, suggesting that these point mutations may operate differently (130). Indeed, these four listed point mutations are scattered across the second, third, and fourth blades of the six-bladed β-propeller domain.

Varying responses to artemisinin also resulted when the same K13 variant codon was edited in different parasite backgrounds (130). This emphasized the important contribution of the genetic background to artemisinin resistance and the presence of secondary determinants that can modulate the degree of resistance. Candidate secondary loci suggested to modulate the artemisinin response were identified in various GWASs (22, 89, 135). Miotto et al. (89) identified the MDR2 (multidrug resistance protein 2) T484I, ferredoxin D193Y, PIB7 (phosphoinositide-binding protein 7) C1484F, ARPS10 (apicoplast ribosomal protein S10) V127M, and PfCRT I356T and N326S mutations as components of a genetic founder background in Southeast Asian isolates. However, their contribution to artemisinin resistance has yet to be validated experimentally. Another GWAS of 192 northeast Thai samples identified Kelch10 P623T as a potential modulator of artemisinin resistance via epistatic interactions with the K13 C580Y or E252Q isoforms (20).

K13 mutations that have independently emerged are associated with specific geographical origins (87, 136, 166). While F446I and more recently G533S dominate in Myanmar, and Yunnan, southern China, near the China-Myanmar border at 56% and 44% prevalence respectively, E252Q is found at the western Thai border and in Myanmar (59, 107, 163). R539T and Y493H mutations have been detected in Cambodia but not western Thailand (87). The most prevalent C580Y mutation is widespread in Western Cambodia, Thailand, and Vietnam and across the GMS (87, 158). The prevalence of each mutation is highly dynamic, with some fluctuating over time while others, such as C580Y in Cambodia, trend to fixation (107). The high prevalence of the C580Y mutation is driven by clonal expansion of the DHA- and piperaquine-resistant Pailin lineage. Originating in western Cambodia, this lineage spread to Thailand and Vietnam and then to southern Laos and northeastern Thailand (51, 62). This lineage harbors the PLA1/KEL1 (K13 C580Y and pm23 copy > 1) haplotype identified from microsatellite typing of a region from −31 kb up to +50 kb flanking the k13 gene (63). In vitro comparative growth rate studies have shown significant differences between K13 mutations, with contemporary Cambodian parasites displaying less of a fitness cost than older strains, consistent with beneficial effects of secondary determinants (94, 129).

Recent reports document the emergence of K13 mutations in 2 isolates from Rwanda (P574L and A675V), 14 from Guyana (1.6%; C580Y), and 3 from Papua New Guinea (1.3%; C580Y). These appear to be distinct from the PLA1/KEL1 haplotype (24, 82, 90, 134). An earlier report of an M579I mutation linked to delayed clearance and low-level RSA survival (2.3%) (79) has not been independently confirmed. Moreover, the proportion of nonsynonymous to synonymous SNPs in k13 is significantly lower in Africa than Southeast Asia, suggesting that k13 is not undergoing selection in African countries to the degree observed in Southeast Asia (80). Presumably this is in part a result of the K13-mutant isolates having a fitness cost that is more detrimental to their survival in higher-transmission African settings with a greater prevalence of mixed infections and more cases of untreated infection. K13-mutant isolates from Papua New Guinea and French Guiana, adapted from the field or gene-edited to introduce the C580Y mutation, have exhibited elevated RSA survival rates of 6.8% and 27.6%, respectively (82, 90), showing that mutant K13 is able to impart resistance in those strains. However, there is no information on the clinical PCT1/2 outcome to suggest the emergence of clinically defined K13-mediated resistance to artemisinins in these regions.

3.4. Mechanisms Underpinning K13-Mediated Artemisinin Resistance

Similar to proteins encoded by other drug resistance genes, the K13 protein is essential for the parasite’s intraerythrocytic development, although its level can be reduced up to 50% by regulating its mislocalization using a knock-sideways approach in transgenic parasites (12). K13 has been localized to cytostomes at the parasite periphery as well as to intracellular vesicles that associate with either endocytosis or vesicular trafficking of antigens including PfEMP1 or Rab-mediated protein transport (11, 13, 46, 160). K13 mutations have been thought to mediate artemisinin resistance in rings primarily via the reduced activation of artemisinin drugs and/or an enhanced parasite capacity to remove damaged proteins (Figure 1b). Lowered K13 levels resulting from mutations have been postulated to lead to reduced hemoglobin endocytosis and catabolism in young rings, resulting in lowered levels of free Fe(II)PPIX available to activate artemisinin (13, 160). In vitro edited K13-mutant 3D7 parasites were also more refractory to DHA-induced inhibition of proteasomal function and maintained protein turnover. Rings inherently tolerate >100-fold-higher levels of artemisinin than do the later trophozoite stages, where maximal activity occurs, validating the importance of hemoglobin endocytosis and digestion in artemisinin activation and subsequent cytotoxicity (68, 69). Perturbation of the pathway by disrupting falcipain 2a or adding the cysteine protease inhibitor E64 further demonstrated that reduced Fe(II)PPIX-mediated activation of the drug can lower parasite susceptibility to artemisinin (159). In addition, when artemisinin-resistant K13 mutant parasites were deprived of sufficient external amino acid sources, they were reported to be less fit and failed to developmentally progress compared with K13 wild-type parasites, providing further evidence that K13 mutations impact hemoglobin catabolism and intracellular liberation of globin-derived amino acids (18). K13-mutant field isolates also showed lowered heme levels at the trophozoite stage, and treatment with DHA generated fewer heme adducts, when compared with their isogenic K13 wild-type counterparts (54, 55).

Of note, no change in expression of k13 transcript abundance was detected in K13 mutant clinical isolates (91). In addition, reduced expression of k13 in rings was reported to result in hypersensitivity to artemisinin as measured using 72-h dose-response assays (not RSAs) in piggyBac transposon–generated mutant parasite lines (45, 165), contrary to the expected increased tolerance if there were lower K13 steady-state protein levels. A reduced level of K13 protein was observed in early rings in the Cam3.II R539T mutant but not the Cam3.II C580Y mutant when compared to the isogenic wild-type line (11, 120). Further work is clearly required to evaluate whether K13 protein levels or functionalities are affected across all clinical isolates bearing mutations that associate with clinical resistance to artemisinin.

K13 also alters the cell cycle in K13 C580Y clinical isolates by conferring an extended period of ring-to-trophozoite stage development, which could also account for reduced Fe(II)PPIX-mediated drug activation (58, 91). K13 mutations have been proposed to destabilize the protein and to deregulate the phosphorylation of PfPK4, leading to downstream activation of transcriptional stress response pathways and modulation of growth via phosphorylation of the parasite’s eukaryotic initiation factor 2α (eIF2α) (160, 164). Phosphorylation of eIF2α has also been observed as a stress-sensing regulatory mechanism of reduced growth rates in cells deprived of the amino acid isoleucine (6).

Maintaining protein turnover is vital to cell survival upon treatment with artemisinin, which is known to cause the accumulation of damaged, unfolded, and polyubiquitinated proteins and to inhibit proteasome function (17). K13 mutations have been proposed to remove damaged proteins by upregulating the unfolded protein response (UPR), as shown with field isolates, and lowering the levels of ubiquitinated proteins as part of their enhanced cell stress response (35, 91) (Figure 1b). Several groups have observed synergy between artemisinin and proteasome inhibitors, providing a compelling potential avenue to overcome artemisinin resistance (35, 76, 128, 161). This synergy, however, is present in both K13-mutant and wild-type parasites (128), providing evidence that K13 mutations do not mediate resistance by modulating proteasomal activity. Alternatively, K13 mutations have been proposed to protect against artemisinin’s proteostatic activity by reducing drug binding to and polyubiquitination of PI3K, thereby elevating the levels of PI3P-positive vesicles that engage in cellular processes of protein folding, vesicle-mediated protein export, and the UPR (11, 84).

Recent data also suggest that K13 function is linked to parasite mitochondria. Gnädig et al. (46) observed increased colocalization of K13 with mitochondria following a DHA pulse, especially in K13 mutants. We also observed reversal of mutant-K13-mediated artemisinin resistance upon coincubation with the mitochondrial cytochrome bc1 inhibitor atovaquone (S. Mok, B.H. Stokes, N.F. Gnädig, L.S. Ross, T. Yeo, et al., unpublished data).

3.5. Evidence of Non-k13 Genetic Mediators of Artemisinin Resistance

Several recent studies have shown that artemisinin resistance in P. falciparum can be driven by genes other than k13. Non-K13-mediated resistance in clinical studies has been limited to sporadic reports of delayed parasite clearance following ACT in a handful of clinical isolates in Southeast Asia and Africa (92, 133). On the other hand, in vitro selection studies with the Pikine and Thiès strains from Senegal yielded moderately artemisinin-resistant parasites that had acquired mutations in coronin. This protein contains a WD40 seven-bladed β-propeller domain similar in architecture to K13 (31). CRISPR-Cas9 editing of the coronin R100K, E107V, or G50E mutation into the wild-type parental lines reproduced the heightened RSA to similar levels (5–10%) as selected resistant lines. Separate selection studies with 3D7, another African line, yielded mutations in multiple genes including a stop mutation in the hemoglobinase falcipain 2A; however, genetic validation was not included in that report (110). UBP-1 has also been touted as a candidate artemisinin resistance mediator, based on linkage group selection analysis of a genetic cross in the rodent parasite Plasmodium chabaudi between a selected artemisinin-resistant parent and an artemisinin-sensitive parent (60). Studies with P. falciparum revealed a V3275F mutation in pfubp1 that was associated with resistance in vitro (56); however, this gene was not associated with in vivo resistance in Southeast Asian clinical isolates (61). The S160N mutation in the AP-2 adaptin μ subunit observed in post-ACT-treated African isolates also did not associate with elevated RSA values in vitro (56). None of the genes listed above, however, has been associated with delayed clearance rates of P. falciparum infections following artemisinin treatment in patients. More compelling is the GWAS of isolates from the China-Myanmar border region that found an association between decreased P. falciparum susceptibility to artemisinin derivatives and the T38I mutation in the PI3P-binding autophagy-related protein ATG18 (149). It will be interesting to determine whether these genes or others might act in the same functional pathway as k13 and produce similar resistance levels independent of K13 mutations.

3.6. Features of Artemisinin-Induced Dormancy and Recrudescence

In bacterial systems, dormancy, i.e., the process where cells enter a low-energy resting state, has been utilized as a bet-hedging strategy to overcome unfavorable environmental conditions(65). Growth reinitiates once the cells sense that growth conditions have improved (39). This phenomenon has been observed in Plasmodium parasites following short-term treatment (6 to 144 h) with antimalarial compounds including artemisinin derivatives, atovaquone, proguanil, pyrimethamine, and mefloquine or exposure to stress-inducing conditions including cold shock or nutrient starvation (57, 71, 88, 95, 137, 138). Dormant parasites, characterized by condensed chromatin and reduced cytoplasm, are typically resistant to perturbations, suggesting that dormancy is a cell-adapted response to external stress (9). As an example, DHA treatment of P. falciparum parasites led to a temporary developmental arrest followed 9 to 20 days later by recrudescence of a small proportion (0.04–1.3%) of parasites (137). Interestingly, only asexual rings transitioned into dormant forms. Modeling studies have suggested that artemisinin-induced dormancy contributes to recrudescence of artemisinin-resistant parasites and treatment failure (25).

Artemisinin-treated dormant P. falciparum parasites were observed to have reduced metabolic levels yet maintained active apicoplasts and mitochondria (23, 104). Interestingly, atovaquone delayed the recovery of DHA-induced dormant parasites (104), suggesting that mitochondrial activity is critical for survival and regrowth of dormant parasites. Several studies have proposed mechanisms that underpin these processes. One study linked P. falciparum asexual blood stage cell cycle progression with the regulated expression of mitotic Ca2+-dependent kinases (cdpk4, pk2, nima, and ark2) and the prereplicative complex (141). In another study, recovery from DHA-induced dormancy was associated with altered expression levels of cyclins (pfcrk1, pfcrk4) and cyclin-dependent kinases (50). Recovery following isoleucine starvation was also associated with stress-response GCN2-mediated eIF2α phosphorylation, higher transcript levels of the RNA polymerase III repressor PfMaf1, and increased pre-tRNA expression to maintain translation of proteins (6, 85).

Over time, repeated drug pressures and induction of dormancy in wild-type parasite populations might progressively select for faster-recovering rings and lead to artemisinin-resistant parasites that no longer require dormancy, as postulated in a two-step process of artemisinin resistance acquisition (25). This would lead to longer times required for artemisinin treatment to clear parasites and result in higher rates of recrudescence. An in vitro example of this scenario may explain the establishment of an artemisinin-resistant K13 M476I mutant following multiple rounds of short-term artemisinin pressure and subsequent parasite recovery (4, 156). This mutant line was observed to form quiescent, nonpyknotic rings that could rapidly resume intraerythrocytic development upon removing artemisinin pressure, suggesting that mutant K13 obviated the need for parasites to enter dormancy as a survival mechanism (156).

4. CONCLUSIONS

The mechanisms of resistance to compounds that target parasite DV processes, such as hemoglobin degradation and hemozoin formation, are primarily linked to mutations in the transmembrane proteins PfCRT and PfMDR1. As mentioned above, these proteins transport drugs either away from or toward their primary site of action. Resistance to the endoperoxide artemisinins, mediated by mutations in K13, has been associated with reduced activation of the drug and an enhanced UPR. Sets of mutations or specific point mutations can confer a variety of fitness costs that together with the level of drug susceptibility under the relevant drug pressure influence which haplotypes dominate in field isolates. Combining suitable sets of drugs that apply opposing selective pressures in combination therapies provides important opportunities for mitigating drug resistance. This goal can only be achieved with a comprehensive understanding of the mechanisms by which parasites evade various drug pressures.

Supplementary Material

Supplemental Table 1

SUMMARY POINTS.

  1. Progress made in reducing malaria cases and fatalities worldwide is threatened by the emergence of multidrug-resistant P. falciparum parasites.

  2. PfCRT is an essential drug/metabolite transporter, present on the parasite’s DV membrane, that harbors an electronegative drug-binding cavity that presumably transports hemoglobin-derived peptides and that harbors unique, geographically dependent sets of mutations. These mutations can give rise to parasite resistance to chloroquine and in some cases amodiaquine. Resistance to piperaquine has also emerged in Southeast Asia through the gain of individual mutations that in many cases can result in loss of chloroquine resistance. These mutations confer fitness costs that influence their relative abundance within parasite populations.

  3. Certain mutant PfCRT isoforms efficiently transport positively charged drugs, including chloroquine, amodiaquine, or piperaquine, that accumulate in the acidic DV and that prevent heme detoxification.

  4. The other DV transporter, PfMDR1, modulates parasite susceptibility to certain heme-binding antimalarials including chloroquine, amodiaquine, lumefantrine, and mefloquine via copy number variations or point mutations.

  5. Artemisinin derivatives are activated via Fe(II)PPIX-dependent cleavage of their endoperoxide bridge, thereby catalyzing the generation of cytotoxic reactive oxygen species. The activated drug affects multiple cellular processes by alkylating proteins, lipids, and heme.

  6. Single point mutations in K13 constitute the primary genetic determinant of in vitro and in vivo artemisinin resistance. The C580Y mutation appears to be relatively fit in Southeast Asian parasites and is highly prevalent across the GMS.

  7. Secondary genetic determinants in epistasis with k13 appear to modulate the degree of artemisinin resistance.

  8. K13 mutations have been proposed to reduce the protein’s function and to mediate resistance by reducing artemisinin activation, prolonging the quiescent ring stage, upregulating the UPR, and augmenting the elimination of damaged proteins.

FUTURE ISSUES.

  1. Will K13 mutations begin to increase in prevalence in Africa, where malaria causes by far the most deaths, and compromise artemisinin efficacy? Is South America also at risk?

  2. Will lumefantrine resistance arise in Africa? Can we predict the genetic basis and therefore generate candidate molecular markers for epidemiological screening and advance warning?

  3. Can we combine new and past antimalarials into double or triple combination therapies to mitigate the spread of resistance by conferring opposing selective pressures on parasite populations?

  4. Which approaches will be the most effective in identifying drug targets of antimalarials with complex modes of action such as artemisinin, as well as resistance mechanisms and markers, in order to gain a deeper insight into possibilities for overcoming drug resistance?

  5. Experimental elucidation of molecular markers of emerging resistance to ACT drugs is empowered by genetic and genomic approaches that include the use of genome-wide association studies, genetic crosses between P. falciparum drug-resistant and drug-sensitive parasites in humanized mice, and CRISPR/Cas9-based gene editing.

  6. New antimalarials with novel modes of action, or those that inhibit the transporters responsible for resistance, are urgently needed to replace artemisinins and current artemisinin-based combination therapies if or when they fail.

ACKNOWLEDGMENTS

S.M. is grateful to receive support from a long-term fellowship funded by the Human Frontier Science Program Organization (LT000976/2016-L). D.A.F. gratefully acknowledges support from the NIH (R37 AI05234, R01 AI124678, R01 AI109023, R01 AI147628), the Bill & Melinda Gates Foundation (OPP1201387), the Department of Defense (W81XWH1910086), and the Medicines for Malaria Venture (MMV08/0015).

DISCLOSURE STATEMENT

The authors are not aware of any affiliations, memberships, funding, or financial holdings that might be perceived as affecting the objectivity of this review.

Glossary

ACT

artemisinin-based combination therapy

DHA

dihydroartemisinin

GMS

Greater Mekong Subregion

DV

digestive vacuole

Fe(II)PPIX

iron protoporphyrin IX

PfCRT

Plasmodium falciparum chloroquine resistance transporter

Cryo-EM

cryogenic electron microscopy

PfMDR1

Plasmodium falciparum multidrug resistance 1 transporter

SNP

single-nucleotide polymorphism

GWAS

genome-wide association study

pm23

plasmepsin 2–3

PCT1/2 (parasite clearance half-life)

the time in hours that it takes for parasitemia to decrease by 50%

RSA

ring stage survival assay

UPR

unfolded protein response

Footnotes

NOTE ADDED IN PROOF

Recent data from Rwanda show the emergence of K13 R561H mutants in P. falciparum parasites of local origin, and this mutation was shown to confer artemisinin resistance in gene-edited Dd2 parasites (139a).

LITERATURE CITED

  • 1.Agrawal S, Moser KA, Morton L, Cummings MP, Parihar A, et al. 2017. Association of a novel mutation in the Plasmodium falciparum chloroquine resistance transporter with decreased piperaquine sensitivity. J. Infect. Dis 216:468–76 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Amaratunga C, Lim P, Suon S, Sreng S, Mao S, et al. 2016. Dihydroartemisinin-piperaquine resistance in Plasmodium falciparum malaria in Cambodia: a multisite prospective cohort study. Lancet Infect. Dis 16:357–65 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Amato R, Lim P, Miotto O, Amaratunga C, Dek D, et al. 2017. Genetic markers associated with dihydroartemisinin-piperaquine failure in Plasmodium falciparum malaria in Cambodia: a genotype-phenotype association study. Lancet Infect. Dis 17:164–73 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Ariey F, Witkowski B, Amaratunga C, Beghain J, Langlois AC, et al. 2014. A molecular marker of artemisinin-resistant Plasmodium falciparum malaria. Nature 505:50–55 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Ashley EA, Dhorda M, Fairhurst RM, Amaratunga C, Lim P, et al. 2014. Spread of artemisinin resistance in Plasmodium falciparum malaria. N. Engl. J. Med 371:411–23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Babbitt SE, Altenhofen L, Cobbold SA, Istvan ES, Fennell C, et al. 2012. Plasmodium falciparum responds to amino acid starvation by entering into a hibernatory state. PNAS 109:E3278–87 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Bakouh N, Bellanca S, Nyboer B, Moliner Cubel S, Karim Z, et al. 2017. Iron is a substrate of the Plasmodium falciparum chloroquine resistance transporter PfCRT in Xenopus oocytes. J. Biol. Chem 292:16109–21 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Baro NK, Pooput C, Roepe PD. 2011. Analysis of chloroquine resistance transporter (CRT) isoforms and orthologues in S. cerevisiae yeast. Biochemistry 50:6701–10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Barrett MP, Kyle DE, Sibley LD, Radke JB, Tarleton RL. 2019. Protozoan persister-like cells and drug treatment failure. Nat. Rev. Microbiol 17:607–20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Bellanca S, Summers RL, Meyrath M, Dave A, Nash MN, et al. 2014. Multiple drugs compete for transport via the Plasmodium falciparum chloroquine resistance transporter at distinct but interdependent sites. J. Biol. Chem 289:36336–51 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Bhattacharjee S, Coppens I, Mbengue A, Suresh N, Ghorbal M, et al. 2018. Remodeling of the malaria parasite and host human red cell by vesicle amplification that induces artemisinin resistance. Blood 131:1234–47 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Birnbaum J, Flemming S, Reichard N, Soares AB, Mesen-Ramirez P, et al. 2017. A genetic system to study Plasmodium falciparum protein function. Nat. Methods 14:450–56 [DOI] [PubMed] [Google Scholar]
  • 13.Birnbaum J, Scharf S, Schmidt S, Jonscher E, Hoeijmakers WAM, et al. 2020. A Kelch13-defined endocytosis pathway mediates artemisinin resistance in malaria parasites. Science 367:51–59 [DOI] [PubMed] [Google Scholar]
  • 14.Blasco B, Leroy D, Fidock DA. 2017. Antimalarial drug resistance: linking Plasmodium falciparum parasite biology to the clinic. Nat. Med 23:917–28 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Bopp S, Magistrado P, Wong W, Schaffner SF, Mukherjee A, et al. 2018. Plasmepsin II-III copy number accounts for bimodal piperaquine resistance among Cambodian Plasmodium falciparum. Nat. Commun 9:1769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Bray PG, Mungthin M, Hastings IM, Biagini GA, Saidu DK, et al. 2006. PfCRT and the trans-vacuolar proton electrochemical gradient: regulating the access of chloroquine to ferriprotoporphyrin IX. Mol. Microbiol 62:238–51 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Bridgford JL, Xie SC, Cobbold SA, Pasaje CFA, Herrmann S, et al. 2018. Artemisinin kills malaria parasites by damaging proteins and inhibiting the proteasome. Nat. Commun 9:3801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Bunditvorapoom D, Kochakarn T, Kotanan N, Modchang C, Kumpornsin K, et al. 2018. Fitness loss under amino acid starvation in artemisinin-resistant Plasmodium falciparum isolates from Cambodia. Sci. Rep 8:12622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Callaghan PS, Hassett MR, Roepe PD. 2015. Functional comparison of 45 naturally occurring isoforms of the Plasmodium falciparum chloroquine resistance transporter (PfCRT). Biochemistry 54:5083–94 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Cerqueira GC, Cheeseman IH, Schaffner SF, Nair S, McDew-White M, et al. 2017. Longitudinal genomic surveillance of Plasmodium falciparum malaria parasites reveals complex genomic architecture of emerging artemisinin resistance. Genome Biol. 18:78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Chaorattanakawee S, Saunders DL, Sea D, Chanarat N, Yingyuen K, et al. 2015. Ex vivo drug susceptibility testing and molecular profiling of clinical Plasmodium falciparum isolates from Cambodia from 2008 to 2013 suggest emerging piperaquine resistance. Antimicrob. Agents Chemother 59:4631–43 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Cheeseman IH, Miller BA, Nair S, Nkhoma S, Tan A, et al. 2012. A major genome region underlying artemisinin resistance in malaria. Science 336:79–82 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Chen N, LaCrue AN, Teuscher F, Waters NC, Gatton ML, et al. 2014. Fatty acid synthesis and pyruvate metabolism pathways remain active in dihydroartemisinin-induced dormant ring stages of Plasmodium falciparum. Antimicrob. Agents Chemother 58:4773–81 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Chenet SM, Akinyi Okoth S, Huber CS, Chandrabose J, Lucchi NW, et al. 2016. Independent emergence of the Plasmodium falciparum Kelch propeller domain mutant allele C580Y in Guyana. J. Infect. Dis 213:1472–75 [DOI] [PubMed] [Google Scholar]
  • 25.Cheng Q, Kyle DE, Gatton ML. 2012. Artemisinin resistance in Plasmodium falciparum: a process linked to dormancy? Int. J. Parasitol. Drugs Drug Resist 2:249–55 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Combrinck JM, Fong KY, Gibhard L, Smith PJ, Wright DW, Egan TJ. 2015. Optimization of a multi-well colorimetric assay to determine haem species in Plasmodium falciparum in the presence of antimalarials. Malar. J 14:253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Combrinck JM, Mabotha TE, Ncokazi KK, Ambele MA, Taylor D, et al. 2013. Insights into the role of heme in the mechanism of action of antimalarials. ACS Chem. Biol 8:133–37 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Conrad MD, Rosenthal PJ. 2019. Antimalarial drug resistance in Africa: the calm before the storm? Lancet Infect. Dis 19:e338–51 [DOI] [PubMed] [Google Scholar]
  • 29.Cooper RA, Ferdig MT, Su XZ, Ursos LM, Mu J, et al. 2002. Alternative mutations at position 76 of the vacuolar transmembrane protein PfCRT are associated with chloroquine resistance and unique stereospecific quinine and quinidine responses in Plasmodium falciparum. Mol. Pharmacol 61:35–42 [DOI] [PubMed] [Google Scholar]
  • 30.Cowell A, Winzeler E. 2018. Exploration of the Plasmodium falciparum resistome and druggable genome reveals new mechanisms of drug resistance and antimalarial targets. Microbiol. Insights 11:1178636118808529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Demas AR, Sharma AI, Wong W, Early AM, Redmond S, et al. 2018. Mutations in Plasmodium falciparum actin-binding protein coronin confer reduced artemisinin susceptibility. PNAS 115:12799–804 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Dhingra SK, Redhi D, Combrinck JM, Yeo T, Okombo J, et al. 2017. A variant PfCRT isoform can contribute to Plasmodium falciparum resistance to the first-line partner drug piperaquine. mBio 8:e00303–17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Dhingra SK, Small-Saunders JL, Menard D, Fidock DA. 2019. Plasmodium falciparum resistance to piperaquine driven by PfCRT. Lancet Infect. Dis 19:1168–69 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Dini S, Zaloumis S, Cao P, Price RN, Fowkes FJI, et al. 2018. Investigating the efficacy of triple artemisinin-based combination therapies for treating Plasmodium falciparum malaria patients using mathematical modeling. Antimicrob. Agents Chemother 62:e01068–18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Dogovski C, Xie SC, Burgio G, Bridgford J, Mok S, et al. 2015. Targeting the cell stress response of Plasmodium falciparum to overcome artemisinin resistance. PLOS Biol. 13:e1002132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Dondorp AM, Nosten F, Yi P, Das D, Phyo AP, et al. 2009. Artemisinin resistance in Plasmodium falciparum malaria. N. Engl. J. Med 361:455–67 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Dorn A, Vippagunta SR, Matile H, Jaquet C, Vennerstrom JL, Ridley RG. 1998. An assessment of drug-haematin binding as a mechanism for inhibition of haematin polymerisation by quinoline antimalarials. Biochem. Pharmacol 55:727–36 [DOI] [PubMed] [Google Scholar]
  • 38.Duru V, Khim N, Leang R, Kim S, Domergue A, et al. 2015. Plasmodium falciparum dihydroartemisinin-piperaquine failures in Cambodia are associated with mutant K13 parasites presenting high survival rates in novel piperaquine in vitro assays: retrospective and prospective investigations. BMC Med. 13:305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Dworkin J, Shah IM. 2010. Exit from dormancy in microbial organisms. Nat. Rev. Microbiol 8:890–96 [DOI] [PubMed] [Google Scholar]
  • 40.Fidock DA, Nomura T, Talley AK, Cooper RA, Dzekunov SM, et al. 2000. Mutations in the P. falciparum digestive vacuole transmembrane protein PfCRT and evidence for their role in chloroquine resistance. Mol. Cell 6:861–71 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Frosch AE, Laufer MK, Mathanga DP, Takala-Harrison S, Skarbinski J, et al. 2014. Return of widespread chloroquine-sensitive Plasmodium falciparum to Malawi. J. Infect. Dis 210:1110–14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Gabryszewski SJ, Modchang C, Musset L, Chookajorn T, Fidock DA. 2016. Combinatorial genetic modeling of pfcrt-mediated drug resistance evolution in Plasmodium falciparum. Mol. Biol. Evol 33:1554–70 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Gething PW, Casey DC, Weiss DJ, Bisanzio D, Bhatt S, et al. 2016. Mapping Plasmodium falciparum mortality in Africa between 1990 and 2015. N. Engl. J. Med 375:2435–45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Ghorbal M, Gorman M, Macpherson CR, Martins RM, Scherf A, Lopez-Rubio JJ. 2014. Genome editing in the human malaria parasite Plasmodium falciparum using the CRISPR-Cas9 system. Nat. Biotechnol 32:819–21 [DOI] [PubMed] [Google Scholar]
  • 45.Gibbons J, Button-Simons KA, Adapa SR, Li S, Pietsch M, et al. 2018. Altered expression of K13 disrupts DNA replication and repair in Plasmodium falciparum. BMC Genom. 19:849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Gnädig NF, Stokes BH, Edwards RL, Kalantarov GF, Heimsch KC, et al. 2020. Insights into the intracellular localization, protein associations and artemisinin resistance properties of Plasmodium falciparum K13. PLOS Pathog. 16:e1008482. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Goldberg DE. 2013. Complex nature of malaria parasite hemoglobin degradation. PNAS 110:5283–84. Erratum. 2013. PNAS 110(17):7097 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Goodman CD, Siregar JE, Mollard V, Vega-Rodriguez J, Syafruddin D, et al. 2016. Parasites resistant to the antimalarial atovaquone fail to transmit by mosquitoes. Science 352:349–53 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Gorka AP, de Dios A, Roepe PD. 2013. Quinoline drug-heme interactions and implications for antimalarial cytostatic versus cytocidal activities. J. Med. Chem 56:5231–46 [DOI] [PubMed] [Google Scholar]
  • 50.Gray KA, Gresty KJ, Chen N, Zhang V, Gutteridge CE, et al. 2016. Correlation between cyclin dependent kinases and artemisinin-induced dormancy in Plasmodium falciparum in vitro. PLOS ONE 11:e0157906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Hamilton WL, Amato R, van der Pluijm RW, Jacob CG, Quang HH, et al. 2019. Evolution and expansion of multidrug-resistant malaria in southeast Asia: a genomic epidemiology study. Lancet Infect. Dis 19:943–51 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Hayward R, Saliba KJ, Kirk K. 2005. pfmdr1 mutations associated with chloroquine resistance incur a fitness cost in Plasmodium falciparum. Mol. Microbiol 55:1285–95 [DOI] [PubMed] [Google Scholar]
  • 53.Heinberg A, Kirkman L. 2015. The molecular basis of antifolate resistance in Plasmodium falciparum: looking beyond point mutations. Ann. N. Y. Acad. Sci 1342:10–18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Heller LE, Goggins E, Roepe PD. 2018. Dihydroartemisinin-ferriprotoporphyrin IX adduct abundance in Plasmodium falciparum malarial parasites and the relationship to emerging artemisinin resistance. Biochemistry 57:6935–45 [DOI] [PubMed] [Google Scholar]
  • 55.Heller LE, Roepe PD. 2018. Quantification of free ferriprotoporphyrin IX heme and hemozoin for artemisinin sensitive versus delayed clearance phenotype Plasmodium falciparum malarial parasites. Biochemistry 57:6927–34 [DOI] [PubMed] [Google Scholar]
  • 56.Henrici RC, van Schalkwyk DA, Sutherland CJ. 2019. Modification of pfap2mu and pfubp1 markedly reduces ring-stage susceptibility of Plasmodium falciparum to artemisinin in vitro. Antimicrob. Agents Chemother 64:e01542–19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Henrici RC, van Schalkwyk DA, Sutherland CJ. 2019. Transient temperature fluctuations severely decrease P. falciparum susceptibility to artemisinin in vitro. Int. J. Parasitol. Drugs Drug Resist 9:23–26 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Hott A, Casandra D, Sparks KN, Morton LC, Castanares GG, et al. 2015. Artemisinin-resistant Plasmodium falciparum parasites exhibit altered patterns of development in infected erythrocytes. Antimicrob. Agents Chemother 59:3156–67 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Huang F, Takala-Harrison S, Jacob CG, Liu H, Sun X, et al. 2015. A single mutation in K13 predominates in Southern China and is associated with delayed clearance of Plasmodium falciparum following artemisinin treatment. J. Infect. Dis 212:1629–35 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hunt P, Afonso A, Creasey A, Culleton R, Sidhu AB, et al. 2007. Gene encoding a deubiquitinating enzyme is mutated in artesunate- and chloroquine-resistant rodent malaria parasites. Mol. Microbiol 65:27–40 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Imwong M, Dondorp AM, Nosten F, Yi P, Mungthin M, et al. 2010. Exploring the contribution of candidate genes to artemisinin resistance in Plasmodium falciparum. Antimicrob. Agents Chemother 54:2886–92 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Imwong M, Hien TT, Thuy-Nhien NT, Dondorp AM, White NJ. 2017. Spread of a single multidrug resistant malaria parasite lineage (PfPailin) to Vietnam. Lancet Infect. Dis 17:1022–23 [DOI] [PubMed] [Google Scholar]
  • 63.Imwong M, Suwannasin K, Kunasol C, Sutawong K, Mayxay M, et al. 2017. The spread of artemisinin-resistant Plasmodium falciparum in the Greater Mekong subregion: a molecular epidemiology observational study. Lancet Infect. Dis 17:491–97 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Ismail HM, Barton V, Phanchana M, Charoensutthivarakul S, Wong MH, et al. 2016. Artemisinin activity-based probes identify multiple molecular targets within the asexual stage of the malaria parasites Plasmodium falciparum 3D7. PNAS 113:2080–85 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Jones SE, Lennon JT. 2010. Dormancy contributes to the maintenance of microbial diversity. PNAS 107:5881–86 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Juge N, Moriyama S, Miyaji T, Kawakami M, Iwai H, et al. 2015. Plasmodium falciparum chloroquine resistance transporter is a H+-coupled polyspecific nutrient and drug exporter. PNAS 112:3356–61 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Kim J, Tan YZ, Wicht KJ, Erramilli SK, Dhingra SK, et al. 2019. Structure and drug resistance of the Plasmodium falciparum transporter PfCRT. Nature 576:315–20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Klonis N, Crespo-Ortiz MP, Bottova I, Abu-Bakar N, Kenny S, et al. 2011. Artemisinin activity against Plasmodium falciparum requires hemoglobin uptake and digestion. PNAS 108:11405–10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Klonis N, Xie SC, McCaw JM, Crespo-Ortiz MP, Zaloumis SG, et al. 2013. Altered temporal response of malaria parasites determines differential sensitivity to artemisinin. PNAS 110:5157–62 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Kuhn Y, Rohrbach P, Lanzer M. 2007. Quantitative pH measurements in Plasmodium falciparum-infected erythrocytes using pHluorin. Cell Microbiol. 9:1004–13 [DOI] [PubMed] [Google Scholar]
  • 71.LaCrue AN, Scheel M, Kennedy K, Kumar N, Kyle DE. 2011. Effects of artesunate on parasite recrudescence and dormancy in the rodent malaria model Plasmodium vinckei. PLOS ONE 6:e26689. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Lee AH, Dhingra SK, Lewis IA, Singh MK, Siriwardana A, et al. 2018. Evidence for regulation of hemoglobin metabolism and intracellular ionic flux by the Plasmodium falciparum chloroquine resistance transporter. Sci. Rep 8:13578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Lehane AM, Hayward R, Saliba KJ, Kirk K. 2008. A verapamil-sensitive chloroquine-associated H+ leak from the digestive vacuole in chloroquine-resistant malaria parasites. J. Cell Sci 121:1624–32 [DOI] [PubMed] [Google Scholar]
  • 74.Lehane AM, van Schalkwyk DA, Valderramos SG, Fidock DA, Kirk K. 2011. Differential drug efflux or accumulation does not explain variation in the chloroquine response of Plasmodium falciparum strains expressing the same isoform of mutant PfCRT. Antimicrob. Agents Chemother 55:2310–18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Lewis IA, Wacker M, Olszewski KL, Cobbold SA, Baska KS, et al. 2014. Metabolic QTL analysis links chloroquine resistance in Plasmodium falciparum to impaired hemoglobin catabolism. PLOS Genet. 10:e1004085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Li H, O’Donoghue AJ, van der Linden WA, Xie SC, Yoo E, et al. 2016. Structure- and function-based design of Plasmodium-selective proteasome inhibitors. Nature 530:233–36 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Li W, Mo W, Shen D, Sun L, Wang J, et al. 2005. Yeast model uncovers dual roles of mitochondria in action of artemisinin. PLOS Genet. 1:e36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Loesbanluechai D, Kotanan N, de Cozar C, Kochakarn T, Ansbro MR, et al. 2019. Overexpression of plasmepsin II and plasmepsin III does not directly cause reduction in Plasmodium falciparum sensitivity to artesunate, chloroquine and piperaquine. Int. J. Parasitol. Drugs Drug Resist 9:16–22 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Lu F, Culleton R, Zhang M, Ramaprasad A, von Seidlein L, et al. 2017. Emergence of indigenous artemisinin-resistant Plasmodium falciparum in Africa. N. Engl. J. Med 376:991–93 [DOI] [PubMed] [Google Scholar]
  • 80.MalariaGen Plasmodium falciparum Community Proj. 2016. Genomic epidemiology of artemisinin resistant malaria. eLife 5:e08714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Martin RE, Marchetti RV, Cowan AI, Howitt SM, Broer S, Kirk K. 2009. Chloroquine transport via the malaria parasite’s chloroquine resistance transporter. Science 325:1680–82 [DOI] [PubMed] [Google Scholar]
  • 82.Mathieu LC, Cox H, Early AM, Mok S, Lazrek Y, et al. 2020. Local emergence in Amazonia of Plasmodium falciparum k13 C580Y mutants associated with in vitro artemisinin resistance. eLife 9:e51015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Maude RJ, Nguon C, Dondorp AM, White LJ, White NJ. 2014. The diminishing returns of atovaquone-proguanil for elimination of Plasmodium falciparum malaria: modelling mass drug administration and treatment. Malar. J 13:380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Mbengue A, Bhattacharjee S, Pandharkar T, Liu H, Estiu G, et al. 2015. A molecular mechanism of artemisinin resistance in Plasmodium falciparum malaria. Nature 520:683–87 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.McLean KJ, Jacobs-Lorena M. 2017. Plasmodium falciparum Maf1 confers survival upon amino acid starvation. mBio 8:e02317–16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Menard D, Dondorp A. 2017. Antimalarial drug resistance: a threat to malaria elimination. Cold Spring Harb. Perspect. Med 7:a025619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Menard D, Khim N, Beghain J, Adegnika AA, Shafiul-Alam M, et al. 2016. A worldwide map of Plasmodium falciparum K13-propeller polymorphisms. N. Engl. J. Med 374:2453–64 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Menard S, Ben Haddou T, Ramadani AP, Ariey F, Iriart X, et al. 2015. Induction of multidrug tolerance in Plasmodium falciparum by extended artemisinin pressure. Emerg. Infect. Dis 21:1733–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Miotto O, Amato R, Ashley EA, MacInnis B, Almagro-Garcia J, et al. 2015. Genetic architecture of artemisinin-resistant Plasmodium falciparum. Nat. Genet 47:226–34 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Miotto O, Sekihara M, Tachibana S-I, Yamauchi M, Pearson RD, et al. 2019. Emergence of artemisinin-resistant Plasmodium falciparum with kelch13 C580Y mutations on the island of New Guinea. bioRxiv 621813 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Mok S, Ashley EA, Ferreira PE, Zhu L, Lin Z, et al. 2015. Population transcriptomics of human malaria parasites reveals the mechanism of artemisinin resistance. Science 347:431–35 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Mukherjee A, Bopp S, Magistrado P, Wong W, Daniels R, et al. 2017. Artemisinin resistance without pfkelch13 mutations in Plasmodium falciparum isolates from Cambodia. Malar. J 16:195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Mukherjee A, Gagnon D, Wirth DF, Richard D. 2018. Inactivation of plasmepsins 2 and 3 sensitizes Plasmodium falciparum to the antimalarial drug piperaquine. Antimicrob. Agents Chemother 62:e02309–17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Nair S, Li X, Arya GA, McDew-White M, Ferrari M, et al. 2018. Fitness costs and the rapid spread of kelch13-C580Y substitutions conferring artemisinin resistance. Antimicrob. Agents Chemother 62:e00605–18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Nakazawa S, Maoka T, Uemura H, Ito Y, Kanbara H. 2002. Malaria parasites giving rise to recrudescence in vitro. Antimicrob. Agents Chemother 46:958–65 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Ncokazi KK, Egan TJ. 2005. A colorimetric high-throughput β-hematin inhibition screening assay for use in the search for antimalarial compounds. Anal. Biochem 338:306–19 [DOI] [PubMed] [Google Scholar]
  • 97.Noedl H, Se Y, Schaecher K, Smith BL, Socheat D, et al. 2008. Evidence of artemisinin-resistant malaria in western Cambodia. N. Engl. J. Med 359:2619–20 [DOI] [PubMed] [Google Scholar]
  • 98.Nsobya SL, Dokomajilar C, Joloba M, Dorsey G, Rosenthal PJ. 2007. Resistance-mediating Plasmodium falciparum pfcrt and pfmdr1 alleles after treatment with artesunate-amodiaquine in Uganda. Antimicrob. Agents Chemother 51:3023–25 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Ochong E, Tumwebaze PK, Byaruhanga O, Greenhouse B, Rosenthal PJ. 2013. Fitness consequences of Plasmodium falciparum pfmdr1 polymorphisms inferred from ex vivo culture of Ugandan parasites. Antimicrob. Agents Chemother 57:4245–51 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Olafson KN, Ketchum MA, Rimer JD, Vekilov PG. 2015. Mechanisms of hematin crystallization and inhibition by the antimalarial drug chloroquine. PNAS 112:4946–51 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Packard RM. 2014. The origins of antimalarial-drug resistance. N. Engl. J. Med 371:397–99 [DOI] [PubMed] [Google Scholar]
  • 102.Parobek CM, Parr JB, Brazeau NF, Lon C, Chaorattanakawee S, et al. 2017. Partner-drug resistance and population substructuring of artemisinin-resistant Plasmodium falciparum in Cambodia. Genome Biol. Evol 9:1673–86 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Patzewitz EM, Salcedo-Sora JE, Wong EH, Sethia S, Stocks PA, et al. 2013. Glutathione transport: a new role for PfCRT in chloroquine resistance. Antioxid. Redox Signal 19:683–95 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Peatey CL, Chavchich M, Chen N, Gresty KJ, Gray KA, et al. 2015. Mitochondrial membrane potential in a small subset of artemisinin-induced dormant Plasmodium falciparum parasites in vitro. J. Infect. Dis 212:426–34 [DOI] [PubMed] [Google Scholar]
  • 105.Pelleau S, Moss EL, Dhingra SK, Volney B, Casteras J, et al. 2015. Adaptive evolution of malaria parasites in French Guiana: Reversal of chloroquine resistance by acquisition of a mutation in pfcrt. PNAS 112:11672–77 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Phuc BQ, Rasmussen C, Duong TT, Dong LT, Loi MA, et al. 2017. Treatment failure of dihydroartemisinin/piperaquine for Plasmodium falciparum malaria, Vietnam. Emerg. Infect. Dis 23:715–17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Phyo AP, Ashley EA, Anderson TJC, Bozdech Z, Carrara VI, et al. 2016. Declining efficacy of artemisinin combination therapy against P. falciparum malaria on the Thai-Myanmar border (2003–2013): the role of parasite genetic factors. Clin. Infect. Dis 63:784–91 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Pulcini S, Staines HM, Lee AH, Shafik SH, Bouyer G, et al. 2015. Mutations in the Plasmodium falciparum chloroquine resistance transporter, PfCRT, enlarge the parasite’s food vacuole and alter drug sensitivities. Sci. Rep 5:14552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Reed MB, Saliba KJ, Caruana SR, Kirk K, Cowman AF. 2000. Pgh1 modulates sensitivity and resistance to multiple antimalarials in Plasmodium falciparum. Nature 403:906–9 [DOI] [PubMed] [Google Scholar]
  • 110.Rocamora F, Zhu L, Liong KY, Dondorp A, Miotto O, et al. 2018. Oxidative stress and protein damage responses mediate artemisinin resistance in malaria parasites. PLOS Pathog. 14:e1006930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Rohrbach P, Sanchez CP, Hayton K, Friedrich O, Patel J, et al. 2006. Genetic linkage of pfmdr1 with food vacuolar solute import in Plasmodium falciparum. EMBO J. 25:3000–11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Roper C, Pearce R, Nair S, Sharp B, Nosten F, Anderson T. 2004. Intercontinental spread of pyrimethamine-resistant malaria. Science 305:1124. [DOI] [PubMed] [Google Scholar]
  • 113.Rosenthal PJ. 2013. The interplay between drug resistance and fitness in malaria parasites. Mol. Microbiol 89:1025–38 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Ross LS, Dhingra SK, Mok S, Yeo T, Wicht KJ, et al. 2018. Emerging Southeast Asian PfCRT mutations confer Plasmodium falciparum resistance to the first-line antimalarial piperaquine. Nat. Commun 9:3314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Ross LS, Fidock DA. 2019. Elucidating mechanisms of drug-resistant Plasmodium falciparum. Cell Host Microbe 26:35–47 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Sa JM, Twu O, Hayton K, Reyes S, Fay MP, et al. 2009. Geographic patterns of Plasmodium falciparum drug resistance distinguished by differential responses to amodiaquine and chloroquine. PNAS 106:18883–89 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Sanchez CP, Rohrbach P, McLean JE, Fidock DA, Stein WD, Lanzer M. 2007. Differences in trans-stimulated chloroquine efflux kinetics are linked to PfCRT in Plasmodium falciparum. Mol. Microbiol 64:407–20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.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 70:786–98 [DOI] [PubMed] [Google Scholar]
  • 119.Shaw PJ, Chaotheing S, Kaewprommal P, Piriyapongsa J, Wongsombat C, et al. 2015. Plasmodium parasites mount an arrest response to dihydroartemisinin, as revealed by whole transcriptome shotgun sequencing (RNA-seq) and microarray study. BMC Genom. 16:830. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Siddiqui G, Srivastava A, Russell AS, Creek DJ. 2017. Multi-omics based identification of specific biochemical changes associated with PfKelch13-mutant artemisinin-resistant Plasmodium falciparum. J. Infect. Dis 215:1435–44 [DOI] [PubMed] [Google Scholar]
  • 121.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 194:528–35 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Sidhu AB, Verdier-Pinard D, Fidock DA. 2002. Chloroquine resistance in Plasmodium falciparum malaria parasites conferred by pfcrt mutations. Science 298:210–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Skinner-Adams TS, Fisher GM, Riches AG, Hutt OE, Jarvis KE, et al. 2019. Cyclization-blocked proguanil as a strategy to improve the antimalarial activity of atovaquone. Commun. Biol 2:166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Small-Saunders JL, Hagenah L, Fidock DA. 2020. Turning the tide: targeting PfCRT to combat drug-resistant P. falciparum? Nat. Rev. Microbiol 18:261–62 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Spring MD, Lin JT, Manning JE, Vanachayangkul P, Somethy S, et al. 2015. Dihydroartemisinin-piperaquine failure associated with a triple mutant including kelch13 C580Y in Cambodia: an observational cohort study. Lancet Infect. Dis 15:683–91 [DOI] [PubMed] [Google Scholar]
  • 126.St. Laurent B, Miller B, Burton TA, Amaratunga C, Men S, et al. 2015. Artemisinin-resistant Plasmodium falciparum clinical isolates can infect diverse mosquito vectors of Southeast Asia and Africa. Nat. Commun 6:8614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Staines HM, Burrow R, Teo BH, Chis Ster I, Kremsner PG, Krishna S. 2018. Clinical implications of Plasmodium resistance to atovaquone/proguanil: a systematic review and meta-analysis. J. Antimicrob. Chemother 73:581–95 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Stokes BH, Yoo E, Murithi JM, Luth MR, Afanasyev P, et al. 2019. Covalent Plasmodium falciparum-selective proteasome inhibitors exhibit a low propensity for generating resistance in vitro and synergize with multiple antimalarial agents. PLOS Pathog. 15:e1007722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Straimer J, Gnadig NF, Stokes BH, Ehrenberger M, Crane AA, Fidock DA. 2017. Plasmodium falciparum K13 mutations differentially impact ozonide susceptibility and parasite fitness in vitro. mBio 8:e00172–17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Straimer J, Gnadig NF, Witkowski B, Amaratunga C, Duru V, et al. 2015. K13-propeller mutations confer artemisinin resistance in Plasmodium falciparum clinical isolates. Science 347:428–31 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Sullivan DJ. 2002. Theories on malarial pigment formation and quinoline action. Int. J. Parasitol 32:1645–53 [DOI] [PubMed] [Google Scholar]
  • 132.Sullivan DJ. 2013. Plasmodium drug targets outside the genetic control of the parasite. Curr. Pharm. Des 19:282–89 [PMC free article] [PubMed] [Google Scholar]
  • 133.Sutherland CJ, Lansdell P, Sanders M, Muwanguzi J, van Schalkwyk DA, et al. 2017. pfk13-independent treatment failure in four imported cases of Plasmodium falciparum malaria treated with artemether-lumefantrine in the United Kingdom. Antimicrob. Agents Chemother 61:e02382–16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Tacoli C, Gai PP, Bayingana C, Sifft K, Geus D, et al. 2016. Artemisinin resistance-associated K13 polymorphisms of Plasmodium falciparum in Southern Rwanda, 2010–2015. Am. J. Trop. Med. Hyg 95:1090–93 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Takala-Harrison S, Clark TG, Jacob CG, Cummings MP, Miotto O, et al. 2013. Genetic loci associated with delayed clearance of Plasmodium falciparum following artemisinin treatment in Southeast Asia. PNAS 110:240–45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Takala-Harrison S, Jacob CG, Arze C, Cummings MP, Silva JC, et al. 2015. Independent emergence of artemisinin resistance mutations among Plasmodium falciparum in Southeast Asia. J. Infect. Dis 211:670–79 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Teuscher F, Gatton ML, Chen N, Peters J, Kyle DE, Cheng Q. 2010. Artemisinin-induced dormancy in Plasmodium falciparum: duration, recovery rates, and implications in treatment failure. J. Infect. Dis 202:1362–68 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Thapar MM, Gil JP, Bjorkman A. 2005. In vitro recrudescence of Plasmodium falciparum parasites suppressed to dormant state by atovaquone alone and in combination with proguanil. Trans. R. Soc. Trop. Med. Hyg 99:62–70 [DOI] [PubMed] [Google Scholar]
  • 139.Tu Y. 2016. Artemisinin—a gift from traditional Chinese medicine to the world (Nobel lecture). Angew. Chem. Int. Ed. Engl 55:10210–26 [DOI] [PubMed] [Google Scholar]; 139a. Uwimana A, Legrand E, Stokes BH, Ndikumana J-LM, Warsame M, et al. 2020. Emergence and clonal expansion of in vitro artemisinin-resistant Plasmodium falciparum Kelch13 R561H mutant parasites in Rwanda. Nat. Med. In press [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Valderramos SG, Valderramos JC, Musset L, Purcell LA, Mercereau-Puijalon O, et al. 2010. Identification of a mutant PfCRT-mediated chloroquine tolerance phenotype in Plasmodium falciparum. PLOS Pathog. 6:e1000887. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.van Biljon R, Niemand J, van Wyk R, Clark K, Verlinden B, et al. 2018. Inducing controlled cell cycle arrest and re-entry during asexual proliferation of Plasmodium falciparum malaria parasites. Sci. Rep 8:16581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.van der Pluijm RW, Imwong M, Chau NH, Hoa NT, Thuy-Nhien NT, et al. 2019. Determinants of dihydroartemisinin-piperaquine treatment failure in Plasmodium falciparum malaria in Cambodia, Thailand, and Vietnam: a prospective clinical, pharmacological, and genetic study. Lancet Infect. Dis 19:952–61 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.van der Pluijm RW, Tripura R, Hoglund RM, Pyae Phyo A, Lek D, et al. 2020. Triple artemisinin-based combination therapies versus artemisinin-based combination therapies for uncomplicated Plasmodium falciparum malaria: a multicentre, open-label, randomised clinical trial. Lancet 395:1345–60 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Veiga MI, Dhingra SK, Henrich PP, Straimer J, Gnadig N, et al. 2016. Globally prevalent PfMDR1 mutations modulate Plasmodium falciparum susceptibility to artemisinin-based combination therapies. Nat. Commun 7:11553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Venkatesan M, Gadalla NB, Stepniewska K, Dahal P, Nsanzabana C, et al. 2014. Polymorphisms in Plasmodium falciparum chloroquine resistance transporter and multidrug resistance 1 genes: parasite risk factors that affect treatment outcomes for P. falciparum malaria after artemether-lumefantrine and artesunate-amodiaquine. Am. J. Trop. Med. Hyg 91:833–43 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.von Seidlein L, Peto TJ, Landier J, Nguyen TN, Tripura R, et al. 2019. The impact of targeted malaria elimination with mass drug administrations on falciparum malaria in Southeast Asia: A cluster randomised trial. PLOS Med. 16:e1002745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Wang J, Huang L, Li J, Fan Q, Long Y, et al. 2010. Artemisinin directly targets malarial mitochondria through its specific mitochondrial activation. PLOS ONE 5:e9582. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Wang J, Zhang CJ, Chia WN, Loh CC, Li Z, et al. 2015. Haem-activated promiscuous targeting of artemisinin in Plasmodium falciparum. Nat. Commun 6:10111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Wang Z, Cabrera M, Yang J, Yuan L, Gupta B, et al. 2016. Genome-wide association analysis identifies genetic loci associated with resistance to multiple antimalarials in Plasmodium falciparum from China-Myanmar border. Sci. Rep 6:33891. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Wellems TE, Plowe CV. 2001. Chloroquine-resistant malaria. J. Infect. Dis 184:770–76 [DOI] [PubMed] [Google Scholar]
  • 151.White NJ. 2004. Antimalarial drug resistance. J. Clin. Investig 113:1084–92 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.White NJ. 2011. The parasite clearance curve. Malar. J 10:278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.White NJ, Pukrittayakamee S, Hien TT, Faiz MA, Mokuolu OA, Dondorp AM. 2014. Malaria. Lancet 383:723–35 [DOI] [PubMed] [Google Scholar]
  • 154.Witkowski B, Amaratunga C, Khim N, Sreng S, Chim P, et al. 2013. Novel phenotypic assays for the detection of artemisinin-resistant Plasmodium falciparum malaria in Cambodia: in-vitro and ex-vivo drug-response studies. Lancet Infect. Dis 13:1043–49 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Witkowski B, Duru V, Khim N, Ross LS, Saintpierre B, et al. 2017. A surrogate marker of piperaquine-resistant Plasmodium falciparum malaria: a phenotype-genotype association study. Lancet Infect. Dis 17:174–83 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Witkowski B, Lelievre J, Barragan MJ, Laurent V, Su XZ, et al. 2010. Increased tolerance to artemisinin in Plasmodium falciparum is mediated by a quiescence mechanism. Antimicrob. Agents Chemother 54:1872–77 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.World Health Organ. 2019. World malaria report 2019. Rep., World Health Organ., Geneva. https://www.who.int/publications-detail/world-malaria-report-2019 [Google Scholar]
  • 158.WWARN K13 Genotype-Phenotype Study Group. 2019. Association of mutations in the Plasmodium falciparum Kelch13 gene (Pf3D7_1343700) with parasite clearance rates after artemisinin-based treatments—a WWARN individual patient data meta-analysis. BMC Med. 17:1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Xie SC, Dogovski C, Hanssen E, Chiu F, Yang T, et al. 2016. Haemoglobin degradation underpins the sensitivity of early ring stage Plasmodium falciparum to artemisinins. J. Cell Sci 129:406–16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Yang T, Yeoh LM, Tutor MV, Dixon MW, McMillan PJ, et al. 2019. Decreased K13 abundance reduces hemoglobin catabolism and proteotoxic stress, underpinning artemisinin resistance. Cell Rep. 29:2917–28.e5 [DOI] [PubMed] [Google Scholar]
  • 161.Zhan W, Visone J, Ouellette T, Harris JC, Wang R, et al. 2019. Improvement of asparagine ethylenediamines as anti-malarial Plasmodium-selective proteasome inhibitors. J. Med. Chem 62:6137–45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Zhang H, Howard EM, Roepe PD. 2002. Analysis of the antimalarial drug resistance protein PfCRT expressed in yeast. J. Biol. Chem 277:49767–75 [DOI] [PubMed] [Google Scholar]
  • 163.Zhang J, Li N, Siddiqui FA, Xu S, Geng J, et al. 2019. In vitro susceptibility of Plasmodium falciparum isolates from the China-Myanmar border area to artemisinins and correlation with K13 mutations. Int. J. Parasitol. Drugs Drug Resist 10:20–27 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Zhang M, Gallego-Delgado J, Fernandez-Arias C, Waters NC, Rodriguez A, et al. 2017. Inhibiting the Plasmodium eIF2α kinase PK4 prevents artemisinin-induced latency. Cell Host Microbe 22:766–76.e4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Zhang M, Wang C, Otto TD, Oberstaller J, Liao X, et al. 2018. Uncovering the essential genes of the human malaria parasite Plasmodium falciparum by saturation mutagenesis. Science 360:eaap7847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Zhu L, Tripathi J, Rocamora FM, Miotto O, van der Pluijm R, et al. 2018. The origins of malaria artemisinin resistance defined by a genetic and transcriptomic background. Nat. Commun 9:5158. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental Table 1

RESOURCES