Open access peer-reviewed chapter

Identification and Validation of Novel Drug Targets for the Treatment of Plasmodium falciparum Malaria: New Insights

Written By

Sergey Lunev, Fernando A. Batista, Soraya S. Bosch, Carsten Wrenger and Matthew R. Groves

Submitted: 06 May 2016 Reviewed: 07 September 2016 Published: 30 November 2016

DOI: 10.5772/65659

From the Edited Volume

Current Topics in Malaria

Edited by Alfonso J. Rodriguez-Morales

Chapter metrics overview

3,052 Chapter Downloads

View Full Metrics

Abstract

In order to counter the malarial parasite’s striking ability to rapidly develop drug resistance, a constant supply of novel antimalarial drugs and potential drug targets must be available. The so-called Harlow-Knapp effect, or “searching under the lamp post,” in which scientists tend to further explore only the areas that are already well illuminated, significantly limits the availability of novel drugs and drug targets. This chapter summarizes the pool of electron transport chain (ETC) and carbon metabolism antimalarial targets that have been “under the lamp post” in recent years, as well as suggest a promising new avenue for the validation of novel drug targets. The interplay between the pathways crucial for the parasite, such as pyrimidine biosynthesis, aspartate metabolism, and mitochondrial tricarboxylic acid (TCA) cycle, is described in order to create a “road map” of novel antimalarial avenues.

Keywords

  • malaria
  • Plasmodium falciparum
  • drug design
  • drug target validation
  • protein interference
  • metabolic map
  • oligomerization

1. Introduction

“Portrait of a serial killer,” a commentary published in 2002 in Nature Journal states: “Malaria may have killed half of all the people that ever lived” [1]. Despite the effort and funds spent on malaria eradication, it continues to infect approximately 200 million people worldwide every year and kill one in every four infected [2]. While effective in the past, current antimalarials are becoming less and less reliable as the parasite rapidly develops drug resistance [3]. There have been a number of extensive reviews covering the recent status of antimalarial research and parasite’s resistance [3–11]. The shared message highlighted in these articles is that a constant supply of novel antimalarials is urgently required. Similarly to the Harlow-Knapp effect described for human kinase research [12], the majority of the antimalarial research is currently aimed at optimization of existing drugs targeting the known and validated pathways.

The currently used antimalarial drugs can be classified into few classes based on the mode of action [3, 7]. Briefly, the groups that receive the most attention of the researchers include the artemisinins and chloroquine-like compounds, which target the food vacuole and heme processing and detoxification [13, 14], antifolates targeting the mitochondrial dihydrofolate reductase (DHFR) and dihydropteroate synthase (DHPS), such as proguanil [15, 16], and mitochondrial inhibitors targeting the electron transport chain and consequently the pyrimidine biosynthesis. Unfortunately, resistance has been reported for nearly all available treatments [3, 7]. Unsurprisingly, compounds such as artemisinin and quinolines that target a broad range of essential pathways within the parasite have successfully been used for nearly 40 years before the widespread of resistance had been reported. In contrast, single-target drugs, such as antifolates and atovaquone, have lost their efficacy within few years of clinical use [11, 17]. A number of promising approaches to counter the fast emerging drug resistance suggested by Verlinden et al. include extension of combination therapy to three or more orthogonal drugs, development and use of multitargeting compounds interfering with unrelated targets, and deeper look into the unexplored alternative targets [3]. In all three cases, in order to successfully overcome the parasite’s remarkable ability to develop resistance to nearly all drugs used against it, by far, a number of novel validated drug targets must be significantly expanded.

This chapter summarizes the pool of the mitochondrial and carbon metabolism targets that have been “under the spotlight” in recent years, as well as suggest a promising new avenue for the validation of novel drug targets. We will focus on the interplay between the pathways crucial for the parasite, such as pyrimidine biosynthesis, aspartate metabolism, and mitochondrial TCA cycle, in order to create a “road map” for further antimalarial drug development.

Advertisement

2. The Harlow-Knapp effect

A scientific analogue of biblical “The rich get richer and the poor get poorer” can be rephrased as “the propensity of the biomedical and pharmaceutical research communities to focus their activities, as quantified by the number of publications and patents, on a small fraction of the proteome” [12] or the “Harlow-Knapp effect.” It was first noted by Harlow and colleagues [18] and further expanded by Knapp group [19], based on the analysis of the amount of publications and patents featuring human protein kinases. Kinases are known to regulate the majority of the cellular pathways including those involved in cancer and other diseases. It was observed that despite the availability of human kinome [20] more than three quarters of protein research was still focused on just 10 per cent of the kinases that were already known before the kinome publication [21]. Edwards and co-workers have also noticed that “the availability of research tools influences a protein’s popularity.” In other words, scientists tend to further explore the well-known systems, ignoring the less studied biomolecules where the probing tools are yet unavailable.

The availability of such tools for each system greatly limits the research opportunities and the attention to said system. Antimalarial research is not an exception to Harlow-Knapp effect: a limited opportunity for genetic manipulation [22] and complex life cycle of the parasite makes novel drug target validation highly challenging. Similarly to the human kinase research scientists tend to “keep looking under the spot light” among the few already validated targets, such as mitochondrial bc1 complex in malaria (target of the widely used Atovaquone), trying to optimize the existing compounds. Since first mentioned in the literature, there have been published more than 40 articles featuring plasmodial bc1 complex [23] and to the date it remains one of the most cited plasmodial enzyme.

Dihydroorotate dehydrogenase from Plasmodium falciparum (PfDHODH) is another clear example of the Harlow-Knapp effect in antimalarial research. Since first proposed as a potential drug target more than a decade ago [24] and first inhibitors reported few years later [25], the major part of the research effort was focused on the optimization of the initial scaffold. In addition to the recent achievements in PfDHODH inhibitor discovery by Phillips et al. [26], orthogonal methods, such as fragment-based drug design and virtual screening, have already yielded a number of very potent chemical scaffolds for this enzyme [27].

This divergent approach should be further exploited for other targets in order to yield novel and more potent scaffolds and support the antimalarial research.

Advertisement

3. Combinational therapy

The compound artemisinin and its derivatives have long been considered the most active and potent antimalarials for their efficacy against nearly all parasite stages [9, 14]. Artemisinins are believed to cause alkylation of proteins and heme and lead to oxidative damage within the parasite as well as affect the heme-related detoxification, although the exact mode of action is still a subject of debate [9, 14, 28]. Artemisinin-based combination therapy (ACT) is still recommended by World Health Organisation (WHO) for the treatment of uncomplicated falciparum and non-falciparum malaria in nearly all areas [7]. ACT implies the use of the fast acting artemisinin component, responsible for the rapid parasitemia clearance, in combination with another long-acting drug partner to eliminate the remaining parasites and suppress the selection of artemisinin resistance [29]. Despite the recent widespread of artemisinin-resistant falciparum malaria in Southeast Asia [30], the proven efficacy of combination therapy suggests that there is a pressing need for greater variety of highly effective antimalarial compounds. Combination of two or more drugs with different mode of action and resistance mechanisms significantly lowers the chances of the parasites to develop resistance to such treatment [31]. Thus, the research focus should be extended from optimization of existing compounds to development of novel research tools in order to explore and dissect other potentially druggable pathways of the parasite and thus bypass of the Harlow-Knapp effect. As stated by Verlinden et al.: “History has clearly indicated that new antimalarials must be continually developed in the ensuing event of resistance development to the current antimalarial arsenal.” The occurrence of drug resistance in malaria is significantly faster than the development of antimalarials [3]. Thus, a constant supply of novel unrelated antimalarial compounds with orthogonal modes of action is urgently required.

Advertisement

4. The mitochondria as drug target for P. falciparum malaria

Mitochondria are organelles that act as the power plants of the cell, as they produce energy for all cellular activities. There are several molecular and functional differences between the mitochondria of Plasmodium species and those from the host. It is also known that the plasmodial mitochondria play a critical and essential role in the parasite’s life cycle [5, 32, 33]. Previous studies have suggested that oxidative phosphorylation is not an essential pathway for parasite’s survival during blood stage [34, 35]. In this stage, the parasite depends mainly on glycolysis as an energy source [3638]. The observed glucose consumption in P. falciparum-infected red blood cells (RBC) was 75- to100-fold higher than in uninfected RBC [39]. Extraordinary glucose uptake during the infection leads to hypoglycemia, which together with an increased production of lactate and resulting lactic acidosis, are the major causes of mortality during severe malaria [40]. Thus, it is generally believed that the role of mitochondria in the parasite is not oxidative phosphorylation but the maintenance of the inner mitochondrial potential. Currently, the chemotherapeutic Malarone, a combination of mitochondrial bc1 complex inhibitor Atovaquone and the dihydrofolate reductase inhibitor Proguanil, collapses the inner mitochondrial potential and induces parasite’s growth arrest, confirming the mitochondrial metabolism to be crucial for the viability of the parasite. The importance of mitochondria for Plasmodium development in asexual stage is reinforced by the validation of another component of mitochondrial electron transport chain (ETC), dihydroorotate dehydrogenase (DHODH), as drug target [41, 42].

Advertisement

5. Electron transport chain (ETC)

The plasmodial mitochondrial electron transport chain (ETC) is composed of non-proton motive quinone reductases, such as dihydroorotate dehydrogenase (DHODH), malate-quinone oxidoreductase (MQO), glycerol 3-phosphate dehydrogenase (G3PDH), type II NADH dehydrogenase (NDH2, Alternative Complex I), and succinate dehydrogenase (SDH, Complex II), and proton motive respiratory complexes, including bc1 complex (Complex III), cytochrome c oxidase (Complex IV), and ATP synthase (Complex V) (Figure 1). The ETC requires ubiquinone (coenzyme Q) and cytochrome c1 that function as electron carriers between the complexes [33, 4447]. The (possible) roles of the ETC enzymes and their known inhibitors will be discussed in the following topics.

Figure 1.

Suggested “roadmap” of essential metabolic processes of Plasmodium falciparum such as pyrimidine biosynthesis, aspartate metabolism, and mitochondrial TCA cycle. The map includes already-validated drug targets PfDHODH [24] and cytochrome bc1 complex [23, 43], as well as other promising targets.

5.1. Dihydroorotate dehydrogenase (DHODH)

The P. falciparum enzyme dihydroorotate dehydrogenase (PfDHODH) bridges the ETC and the pyrimidine biosynthesis; PfDHODH catalyzes the key step of oxidation of dihydroorotate to orotate (a precursor for the biosynthesis of pyrimidine bases). The flavin mononucleotide (FMN)-dependent oxidation reaction catalyzed by DHODH can be divided in two half reactions: firstly, the oxidation of dihydroorotate through reduction of FMN and, secondly, the reoxidation of FMNH2 to regenerate the active enzyme. Two electrons resulting from this oxidation reaction are fed into the ETC through Flavin mononucleotide cofactor to ubiquinone, generated at the cytochrome bc1 complex, bridging pyrimidine metabolism and ETC [24, 48, 49]. Inhibition of PfDHODH results in disruption of de novo biosynthesis of pyrimidines [48]. During the blood stage, the parasite depends strictly on this pathway for pyrimidine availability, which is essential for the formation of DNA, RNA, glycoproteins, and phospholipids [44].

Given the essential role of the PfDHODH in the survivability of blood stage parasite and the significant differences to human DHODH [24], it is reasonable that the malarial enzyme has emerged as a novel validated drug target [26, 48, 50]. Inhibition of human DHODH was shown to be effective in treatment of autoimmune diseases, such as rheumatoid arthritis [51, 52]. The development of potent hDHODH inhibitors, such leflunomide and brequinar, led to the search of analogues with potential to inhibit plasmodial DHODH. These analogues were found to be poorly effective [53], potentially due to the differences in leflunomide and brequinar binding sites between human and plasmodial DHODH. These differences make PfDHODH a potential species-specific drug target [24], which was extensively explored by a considerable number of studies. Although early research have not yielded effective results, the following studies have led to important achievements in the discovery of PfDHODH inhibitors, such as benzimidazolyl thiophene-2-carboxamides [5456], s-benzyltriazolopyrimidines [57], N-substituted salicylamides [58], trifluoromethyl phenyl butenamide derivatives [59], and triazolopyrimidine-based inhibitors [25, 6064]. The triazolopyrimidine-based compound DSM265 was shown to be a potent inhibitor of the PfDHODH and Plasmodium vivax DHODH (PvDHODH) with excellent selectivity versus hDHODH [48]. DSM265 has become the first DHODH inhibitor to enter the human antimalarial clinical trials, and preclinical development description was recently published, showing significant differences in DSM265 inhibitory activity between mammalian and plasmodial DHODHs. The kill rate of DSM265 for in vitro blood stage activity has shown to be similar to atovaquone, but significantly lower than observed for artemisinin and chloroquine. In addition, DSM265 has shown favorable pharmacokinetic properties, predicted to provide therapeutic concentrations for more than 8 days after a single oral dose in the range of 200–400 mg, what represents an advantage over current treatment options that are dosed daily. DSM265 was well tolerated in repeat dose, showed cardiovascular safety studies in mice and dogs, was not mutagenic, and was inactive against panels of human enzymes/receptors. Together, these data suggest that DSM265 has a high potential to be validated as a drug combination partner for either single-dose treatment or once-weekly chemoprevention [26].

5.2. Cytochrome bc1 (complex III)

The cytochrome bc1, also known as ubiquinol:cytochrome c oxidoreductase or complex III, is the only enzyme complex common to almost all respiratory ETCs [65]. This complex is composed of 11 different polypeptides, and its catalytic core is composed of three subunits, namely cytochrome b, cytochrome c1, and Rieske protein, also known as iron-sulfur protein (ISP) [6668]. Cytochrome bc1 is found in the inner mitochondrial membrane and functions as a transporter of protons into the intermembrane space through the oxidation and reduction of ubiquinone in the Q cycle [6770]. This enzymatic complex contains two distinct binding sites for the reduction and oxidation of ubiquinol and ubiquinone, both located within cytochrome b. The Qo site acts to oxidize ubiquinol near the intermembrane space, whereas the Qi site binds and reduces ubiquinone near the mitochondrial matrix [71, 72].

Although the crystal structure of plasmodial bc1 complex has not been solved, the high degree of sequence homology with other organisms of which the X-ray crystal structure is known (e.g. Saccharomyces cerevisiae [73]), allowed the discovery of many inhibitors. Cytochrome bc1 of Plasmodium is in fact a major drug target for the treatment and prevention of malaria and, to date, is the only component of the ETC with a clinically used antimalarial drug association [23, 43]. The compound atovaquone, a hydroxynaphthoquinone, inhibits cytochrome bc1 by binding to the Qo site. This inhibition leads to parasite death through the collapse of the Plasmodium mitochondrial membrane potential with no effect on the mammalian host [42, 74, 75]. Although atovaquone is a potent plasmodial bc1 complex inhibitor, its clinical utility is limited by the rapid emergence of resistant parasites when used as monotherapy [76]. Resistance to atovaquone has been developed due to mutations in the codon 268 (Y268S/C/N). These mutations affect the binding of the atovaquone to the target [77]. Because of that, atovaquone is used together with proguanil (Malarone) for treating uncomplicated malaria or as chemoprophylaxis for preventing malaria in travellers.

Aside of atovaquone, other bc1 complex inhibitors were described, as acridones [78], quinolones [7981], pyridones [82, 83], and benzene sulfonamides [84]. Although many compounds have presented inhibitory potential against bc1 complex, this target might be considered underexploited, since the majority of these compounds target the Qo site [85]. The Qi site of cytochrome bc1 has been far less explored and only the binding of a few compounds has been reported [8689].

5.3. Type II NADH dehydrogenase (NDH2)

Instead of the canonical multimeric complex I, or NADH:dehydrogenase, found in mammalian mitochondria, the Plasmodium ETC possesses the type II NADH:quinone oxidoreductase (NDH2). This enzyme, also known as alternative complex I, is a five quinone-dependent oxidoreductase enzyme involved in the redox reaction of NADH oxidation with subsequent quinol production [90]. Although the activity of NDH2 is still not biochemically confirmed in P. falciparum, it has been described in some detail for other organisms that also possess the type II NADH:quinone oxidoreductase, such as plants, fungi, and bacteria [9196]. Differently from complex I, NDH2 is not involved in the direct pumping of protons across the membrane. Instead of proton pumping, NDH2 enables the H+-unregulated generation of mitochondrial reducing power supplying the various respiratory chains with reducing equivalents from NAD(P)H [45, 90].

So far, no crystal structure of the P. falciparum NDH2 (PfNDH2) is available, and prediction of PfNDH2 is based on sequence and structural similarities to other redox enzymes [45, 91, 97]. Although reverse genetics of PfNDH2 was shown to be not lethal [98], PfNDH2 was described as a putative “choke point” in the mitochondrial ETC and has been highlighted as a potential target for antimalarial development [45, 90, 99]. Given the lack of structural data for PfNDH2 and its poor homology to any other structure in PDB, the existing studies aiming to inhibit PfNDH2 for “druggable” proposes have used chemoinformatics and virtual screening methods. PfNDH2 (as other NDH2 analogues) has shown to be insensitive to rotenone, a well-known inhibitor of complex I [90, 100]. The compound 1-hydroxy-2-dodecyl-4(1H)quinolone (HDQ), initially identified as an inhibitor of yeast NDH2 [101], was reported to be a potent inhibitor of P. falciparum proliferation [102]. In fact, HDQ inhibits PfNDH2 but, in addition, it disrupts mitochondrial function through the potent inhibition of the bc1 complex [103]. The compounds dibenziodolium chloride (DPI) and diphenyliodonium chloride (IDP) have also been reported to inhibit PfNDH2 activity in crude lysate fractions and both have shown efficacy against whole parasite proliferation [90]. However, a further study put the potential of PfNDH2 inhibition by these compounds into question, since the authors were unable to corroborate the previous findings through dose-effect profiles using purified recombinant PfNDH2 [100]. These results suggest that DPI and IDP may not be effective inhibitors of PfNDH2, but their antiparasitic effect might be attributed to other enzymes instead (e.g. PfDHODH) [100]. Inhibition of PfNDH2 by artemisinin has also been demonstrated, suggesting a dual role for mitochondria in the action of artemisinin [104]. More recently, Antoine et al. [105] demonstrated that the low degree of inhibition of this enzyme by artemisinin indicates a non-ETC mode of action.

In more recent efforts, Biagini et al. [81] undertook a high-throughput screen (HTS) against PfNDH2 using HDQ in combination with a range of chemoinformatics as starting point. This approach led to the selection of the quinolone core as the key target for SAR, followed by the selection of CK-2-68 as a lead for further development [81, 106]. Structural alterations aiming to improve the inhibitory activity and aqueous solubility led to the compounds SL-2-64 and SL-2-25, the last presenting activity against PfNDH2 and whole-cell P. falciparum at nanomolar range. In vivo experiments using Plasmodium berghei-infected mice demonstrated that SL-2-25 was able to clear parasitemia in the Peters’ standard 4-day suppressive test when given orally a dose of 20 mg kg−1 [107]. SL-2-25, as other quinolones in this study, had the ability to inhibit both PfNDH2 and cytochrome bc1 at low nanomolar range, the same dual inhibition previously observed for HDQ. This dual targeting of two key mitochondrial enzymes suggests that the quinolone pharmacophore is a privileged scaffold for inhibition of both drug targets.

Although the recent efforts to inhibit NDH2 with antimalarial purposes have been a good improvement in the knowledge of its potential as a drug target, the report of PfNDH2 crystal structure would allow a deep investigation on both biochemical characterization and drug design targeting PfNDH2.

5.4. Mitochondrial glycerol-3-phosphate dehydrogenase (mG3DH)

Mitochondrial glycerol 3-phosphate dehydrogenase (mG3DH) is a ubiquinone-linked flavoprotein embedded in the mitochondrial inner membrane that transfers reducing equivalents directly from glycerol 3-phosphate into the electron transport chain [108, 109]. The P. falciparum genome has homologues of both cytoplasmic and mitochondrial G3DH and assays indicate that the addition of glycerol-3-phosphate stimulates electron transport through the inner membrane [110112]. Together with NDH2, mitochondrial G3DH from P. falciparum (PfmG3DH) is also suggested to play an important role in the redox balance under conditions of low O2. Further studies might clarify the essentiality of mG3PDH in Plasmodium survivability and also evaluate its potential as a drug target.

5.5. Succinate dehydrogenase (SDH)

The succinate dehydrogenase (SDH), also known as succinate: ubiquinone oxidoreductase (SQO) or complex II, is an enzymatic complex involved in both TCA cycle, functioning as a primary dehydrogenase, and in mitochondrial ETC, functioning as electron donator [113]. This dual role makes SDH a direct connection between major systems in aerobic energy metabolism. The enzyme has been isolated and characterized from prokaryotic [114117] and eukaryotic organisms [118121], including P. falciparum [122, 123]. SDH is located in the cytoplasmic membrane in bacteria [124] and in the mitochondrial inner membrane in eukaryotes [125]. The enzymatic complex is highly conserved and is basically composed of four subunits: a flavoprotein subunit (SDH1) and an iron-sulfur subunit (SDH2) together form a soluble heterodimer that binds to a membrane anchor b-type cytochrome (a CybL (SDH3)/CybS (SDH4) heterodimer). In P. falciparum, the two major subunits possess molecular masses of 55 kDa (Fp, flavoprotein subunit) and 35 kDa (Ip, iron-sulfur protein subunit) [122]. The SDH activity has shown to be essential for Plasmodium survivability, what makes this enzyme an attractive target for antimalarial development. The already reported differences in kinetic properties between P. falciparum SDH (PfSDH) and human SDH increase the probability that PfSDH inhibitors might represent potent and selective antimalarial compounds [122]. In fact, SDH has shown sensitivity to a number of inhibitors, such as 5-substituted 2,3-dimethoxy-6-phytyl-1,4-benzoquinone derivatives, plumbagin and licochalcone [125], but so far, inhibitors with potential for antimalarial development still have to be discovered.

5.6. Malate-quinone oxyreductase (MQO)

The malate-quinone oxidoreductase (MQO) is a peripheral membrane-bound flavoprotein, which catalyzes the oxidation of malate to oxaloacetate, reducing ubiquinone [126]. Plasmodium species possesses a group 2 MQO, in contrast to bacterial group 1 MQO [127]. P. falciparum MQO (PfMQO) is part of both mitochondrial ETC and TCA cycle, substituting other mitochondrial malate dehydrogenases (MDH) [111, 112, 128]. To date, no crystal structure of the Plasmodium MQO or inhibition studies are available. However, recent experiments showed that while knockout of six enzymes of plasmodial TCA cycle did not cause any significant growth inhibition, no viable MQO-knockout strains of P. falciparum could be obtained yet [34]. These findings as well as the absence of MQO in the human host make the enzyme an interesting target for antimalarial drug discovery.

5.7. ATPase

Although malaria parasites generate most of their ATP through aerobic glycolysis during the blood stage of their life cycle, they appear to possess a complete ATP synthase complex [47]. P. falciparum ATP synthase (PfATP synthase) is not reported to generate ATP but is suggested to act as a proton leak for the ETC [46, 47]. The use of bedaquiline, TMC207, has been proven to be effective for the treatment of multidrug-resistant tuberculosis. This compound targets Mycobacterium tuberculosis ETC through inhibition of ATP synthase rising the hypothesis that this may also be a valid drug target for malaria in the future [129]. So far, only one PfATP synthase inhibitor was described. The compound almitrine, originally developed as a respiratory stimulant, has activity against PfATP synthase and at the cellular level [130]. Recently, a genetic study demonstrated that mitochondrial ATP synthase is dispensable in blood stage P. berghei, although is essential in the mosquito phase [131]. For P. falciparum, previous attempts to knock out the mitochondrial ATP synthase subunits were unsuccessful, suggesting an essential role played by this enzyme complex in blood stages of the parasite [47]. The difference in essentiality of ATP synthase between P. falciparum and P. berghei could be explained by a possible distinction in the requirements of the two species for ATP [131]. Still, more studies are needed to define whether or not ATP synthase is essential in P. falciparum blood stage and consequently evaluate its potential as antimalarial target.

Advertisement

6. Tricarboxylic acid (TCA) cycle

While Plasmodium relies mainly on glycolysis during the blood stage, the TCA metabolism does occur in asexual Plasmodium, but at low turnover [35]. The exact function of the plasmodial TCA cycle is still a subject of debate, as it does not seem to function like a conventional TCA cycle. In 2010, a branched TCA pathway has been suggested for the parasite [132] but further retracted [133]. It was proposed that plasmodial TCA enzymes function not only in the classical but also in the reverse direction, generating either reductive or an oxidative pathway, depending on the direction. Both pathways would result in the generation of malate, which is subsequently exported from the mitochondria, with α-ketoglutarate (2OG) being anti-ported to feed both the oxidative and reductive pathways [132]. Depending on the nutrient availability, Plasmodium species might not excrete malate as metabolic waste, utilizing it for metabolic purposes [134].

Further metabolomic studies suggest that P. falciparum utilizes conventional TCA cycle during both sexual and asexual blood stages [35]. Functional respiratory chain appears to be essential for the maintenance of inner mitochondrial membrane potential as well as protein and metabolite transport within the mitochondrion. The authors have also reported an increased sensitivity of gametocyte stages to sodium fluoroacetate (NaFAc). NaFAc was previously reported to inhibit TCA cycle enzyme aconitase in Leishmania [135]. Both sexual and asexual cultures of P. falciparum treated with 1 mM NaFAc showed significant citrate accumulation in the parasite as well as decrease in downstream TCA metabolites, suggesting the specific inhibition of aconitase of P. falciparum. However, no significant growth inhibition of the asexual parasites was observed, while gametocyte development was significantly reduced. These findings provide a potential for future transmission-blocking therapy.

Recently, Ke et al. [34] reported significant flexibility in TCA cycle metabolism of P. falciparum. The knockout experiments with all TCA cycle enzymes showed altered substrate fluxes between mitochondrial and cytosolic pools in nearly all cases. Out of eight enzymes of the TCA cycle, knockout of six enzymes of the TCA cycle showed no detectable growth defects. However, the authors were unable to disrupt the genes encoding fumarate hydratase and malate-quinone oxyreductase, suggesting potentially essential role of these two enzymes in asexual parasite development.

Although the fully functional TCA cycle appears to be dispensable for parasite survival in asexual blood stages [34], the interplay of some TCA enzymes with other essential pathways still represents an interesting target for antimalarial drug development. Below, we describe the role of three enzymes (aspartate aminotransferase, malate dehydrogenase, and fumarate hydratase) in Plasmodium metabolism and also their potential for antimalarial drug discovery. Other enzymes involved in this pathway (e.g., PfSDH, PfMQO) were previously described within the ETC section (see above).

6.1. Aspartate aminotransferase

The enzyme aspartate aminotransferase (AspAT) catalyzes the reversible reaction of l-aspartate and α-ketoglutarate into oxaloacetate and l-glutamate. The AspAT from P. falciparum (PfAspAT) was placed into the Ia subfamily, being the most divergent member of this group. The crystal structure of PfAspAT reveals an architecture similar to that previously determined in the Escherichia coli (1B4X14–17) [136139], yeast cytosolic [140], pig heart cytosolic [141], and mitochondrial and cytosolic chicken [142144] homologues. PfAspAT is a homodimeric enzyme [145, 146], and each subunit consists of a large PLP (cofactor) binding domain, a smaller domain, that shifts the enzyme from “closed” to “open” form in order to provide substrate binding and N-terminal region that stabilizes the interaction between the two monomers into a dimer [142, 147, 148]. Two independent active sites are positioned near the oligomeric interface and are formed by residues from both subunits [146]. The active site is highly conserved between available AspATs, making the design of species-specific inhibitors very challenging. However, it is known that the active site requires the formation of a homodimer, and analysis of AspAT has highlighted the N-terminal region as being highly divergent from other AspAT family members in both sequence and structure [145, 146]. Such a divergence may allow a more specific interference with the parasitic AspAT oligomeric surfaces, which offers a unique opportunity to generate highly specific interference with protein function in vivo. Such an approach will be further discussed in this chapter.

6.2. Malate dehydrogenase

The enzyme malate dehydrogenase (MDH) catalyzes the reversible NAD(P)+-dependent oxidation of oxaloacetate to malate. Like other members of the NAD+-dependent dehydrogenase family, the MDHs possess two functional domains, the catalytic domain and the NAD+-binding domain. Protozoan MDHs are differentiated into two subdivisions: mitochondrial and cytosolic MDHs, the first being part of the TCA cycle, providing oxaloacetate for the generation of citrate and NADH to fuel the mitochondrial electron-transport chain. The mitochondrial MDH is absent in P. falciparum, being replaced by PfMQO (described in ETC section). The cytosolic MDH is present in P. falciparum (PfMDH) acting as a supplier of metabolites, such as malate, to the mitochondria and might be responsible for the generation of reducing equivalents to feed the respiratory chain [149].

The crystal structure of PfMDH has recently been solved [150]. Analysis of PfMDH structure revealed a tetrameric assembly, although isoforms of the enzyme from other species have been reported to be present as either dimers or tetramers. Similar to PfAspAT, the oligomeric nature of PfMDH and the low degree of evolutionally conservation of the oligomeric interface residues provide an opportunity for a highly specific protein interference approach (described further).

6.3. Fumarate hydratase

Fumarate hydratase (FH) is an enzyme that catalyzes the reversible conversion of fumarate to malate. Although P. falciparum contains a fumarate hydratase homologue (PfFH), it differs substantially from the “class II” type enzyme found in yeast and mammalian cells [151, 152]. Instead, the PfFH resembles the iron-sulfur-containing “class I”-type enzymes found in some bacteria and archaea [153]. PfFH was shown to be essential to the asexual stages of the parasite [34]. PfFH was initially suggested to be located within the mitochondrion [153], however, this localization is yet not entirely clear.

Fumarate is a side product of the purine salvage pathway and acts as metabolic intermediate of the TCA cycle. As previously mentioned, P. falciparum does not export fumarate as metabolic waste but converts the metabolite to aspartate through malate and oxaloacetate. Besides, P. falciparum-infected erythrocytes and free parasites incorporate labeled fumarate into the nucleic acid and protein fractions [153]. Taken together, these data provide a biosynthetic function for fumarate hydratase and suggest that this enzyme could therefore be targeted for the development of antimalarial chemotherapeutics.

Advertisement

7. Pyrimidine biosynthetic pathway

A key-step for spreading of malaria parasites in the human host is the extensive and rapid replication of parasite DNA, which depends on the availability of essential metabolites, such as pyrimidines [154, 155]. In the Plasmodium species, besides the DNA, the pyrimidine nucleotide is also involved in the biosynthesis of RNA, phospholipids, and glycoproteins [155157]. Sequencing studies have revealed that, in malaria parasites, the genes encoding for the pyrimidine biosynthetic pathway enzymes have been conserved, whereas those responsible for pyrimidines salvage have not [158]. It means that, while human cells are able to acquire pyrimidines either through de novo synthesis or by salvaging, the malaria parasites lack pyrimidine salvage enzymes and depend exclusively on the de novo pathway as source of pyrimidines for their survival [5, 33]. De novo synthesis from carbamoyl phosphate and aspartic acid follows basically the same steps found in the human host and in other eukaryotes: orotic acid is formed by dihydroorotase (DHOase) and DHODH. The orotic acid is so turned into orotidine 5′-monophosphate (OMP) by addition to 5′-phospo-D-ribosyl-α-1-pyrophosphate, a step carried out by orotate phosphoribosyltransferase (OPRT). OMP is subsequently decarboxylated to uridine 5′-monophosphate (UMP), the precursor of all other pyrimidine nucleotides and deoxynucleotides needed for nucleic acid synthesis [159]. Excepting for PfDHODH, which is discussed in the ETC topic, the enzymes involved in de novo pyrimidine biosynthesis pathway that could potentially be exploited for the discovery of novel antimalarials as discussed below.

7.1. Carbamoyl phosphate synthetase II

Carbamoyl phosphate synthetase II (CPSII) is responsible for the first step of the de novo pyrimidine biosynthesis, catalyzing the formation of carbamoyl phosphate in the cytosol from bicarbonate, glutamine, and ATP [160]. Differently from the human CPSII, CPSII from P. falciparum (PfCPSII) is a monofunctional protein [155]. PfCPSII also differs from its mammalian homologue by the presence of two inserted sequences, located between junctions of the glutamine aminotransferase and synthetase domains [161]. Although the absence of structural information and activity inhibitors, the druggable potential of this enzyme has already been demonstrated by the potent growth inhibitory effect of a synthetic ribozyme with specificity for the PfCPSII gene over P. falciparum cultures [162]. The same synthetic ribozyme has shown no toxicity to mammalian cells. Other mini ribozymes were further redesigned to improve cleavage activities and metabolic stabilities [163]. These results suggest that the discover of compounds capable to inhibit PfCPSII in a specific way might be promising antimalarial candidates, since ribozyme approaches have a significant more challenging application due to target accessibility, stability, specificity, and delivery efficiency [164].

7.2. Aspartate transcarbamoylase (ATC)

Aspartate transcarbamoylase (ATC, EC 2.1.3.2) catalyzes the condensation of aspartate and carbamoyl phosphate to form N-carbamoyl-l-aspartate and inorganic phosphate. Previous studies with human tumor tissues showed significantly elevated levels of ATC nearly in all samples [165]. In P. falciparum, ATC is also present as monofunctional protein, unlike its human homologue. Although a number of publications suggest ATC from P. falciparum to be a promising drug target [166168], it has not been fully characterized and no inhibitors have yet been reported. Recently reported crystal structure of the truncated PfATC revealed high level of sequence conservation among homologous enzymes from other organisms, especially in the active site area [169].

7.3. Dihydroorotase

Similarly to CPSII, P. falciparum dihydroorotase (PfDHOase) is a monofunctional protein and thus differs from the mammalian host, in which the 36.7 kDa enzyme is located on the central part of the 240 kDa CAD multifunctional protein [170]. This enzyme catalyzes the reversible cyclization of N-carbamoyl-l-aspartate (CA-asp) to l-dihydroorotate (L-DHO) [159]. Orotate and a series of 5-substituted derivatives were found to inhibit competitively the purified enzyme from P. falciparum culture. In mice infected with P. berghei, 5-fluoro orotate and 5-amino orotate at a dose of 25 μg/g body weight eliminated parasitemia after a 4-day treatment, an effect comparable to that of the same dose of chloroquine. The infected mice treated with 5-fluoro orotate at a lower dose of 2.5 μg/g had a 95% reduction in parasitemia [171]. The moderate inhibition of PfDHOase by L-6-thiodihydroorotate (TDHO) in cultured parasites induced major accumulation of CP-asp and growth arrest, similar to atovaquone [172]. The analysis of physical, kinetic, and inhibitory properties of the recombinant PfDHOase performed by Krungkrai et al. suggests that specific inhibitors may limit the pyrimidine nucleotide pool in the parasite, but have no significant adverse effect to human host [173]. Although the low amount of information about PfDHOase does not allow to confirm it as a good candidate to antimalarial development, the report of its crystal structure and biochemical characterization could clarify whether this enzyme is essential or not to the parasite’s survivability.

7.4. Orotate phosphoribosyl transferase and orotidine 5′-monophosphate decarboxylase

The last two steps of the pyrimidine biosynthesis in P. falciparum are catalyzed by a heteromeric complex that consists of two homodimers of PfOPRT and PfOPDC encoded by two separate genes [174, 175].

The enzyme orotate phosphoribosyl transferase (OPRT) catalyzes the formation of orotidine 5′-monophosphate (OMP) from α-D-phosphoribosyl pyrophosphate (PRPP) and orotate, the fifth step of the pyrimidine biosynthesis [155]. The OPRT inhibitors reported so far includes the compound 5′-Fluoroorotate, an alternative substrate for this enzyme that was shown to inhibit the in vitro growth of P. falciparum at nanomolar range [176, 177] and to clear parasitemia from P. berghei-infected mice [171]. This antimalarial activity is related to the inactivation of malarial thymidylate synthase by 5′-fluoro-2′-deoxy-UMP metabolite through covalent binding to methylene tetrahydrofolate at the active site. The compound pyrazofurin has also been described as a moderate inhibitor of P. falciparum OPRT (PfOPRT), inhibiting its activity at micromolar range by blocking the maturation of trophozoites to schizonts [176, 178]. Interestingly, pyrazofurin does not affect the OPRT activity in mammalian cells [179].

A recent study of the transition state analogues of PfOPRT also showed that despite the tight binding in vitro, the synthetized compounds failed to inhibit the parasite culture growth in vivo [180182]. No growth inhibition was observed at high compound concentrations up to 100 μM, suggesting poor compound accessibility in vivo.

Recently, the crystal structure of PfOPRT has been reported, which shows a homodimeric assembly, where each of two active sites include amino acids from both chains [183]. Despite the high level of homology with human OPRT, the active site of PfOPRT has few amino acids that differ from HsOPRT. The authors suggest that these differences might lead to the design of selective substrate-like inhibitors in the future.

Orotidine 5′-monophosphate decarboxylase (OPDC) catalyzes the final step of de novo pyrimidine biosynthesis pathway, the decarboxylation of orotidine 5′-monophosphate (OMP) to uridine 5′-monophosphate (UMP), with no need for the presence of a cofactor or metal ion [184]. Many inhibitors of plasmodial OPDC have been described so far, being the most promising inhibitor, the nucleotide 5′-monophosphate analogue xanthosine 5′-monophosphate (XMP) [185]. XMP acts as a competitive inhibitor with tighter binding than OMP. The P. falciparum OPDC inhibition by XMP is highly selective, having a 150-fold preference for the malarial enzyme compared to human OPDC. Other inhibitors include the 6-iodouridine 5′-monophosphate (6-iodo-UMP) [186], 6-azidouridine 5′-monophosphate (6-N3-UMP) [187], barbiturate 5′-monophosphate (BMP) [185], 6-N-methylamino uridine [187], and 6-N,N-dimethylamino uridine [187]. Although a considerable number of PfODC inhibitors have been described and a crystal structure of PfOPDC is available [188], a deeper investigation is necessary to clarify PfOPDC as validated drug target.

Advertisement

8. Protein interference assay (PIA) as drug validation tool

We have recently proposed a novel promising drug-target validation approach that relies on common feature of all biological systems—oligomerization [22]. Oligomerization is a self-assembly of two or more copies of one protein molecule (or different molecules) into one object. Recent analysis shows that majority (60%) of non-redundant protein structures available in the Protein Data Bank (PDB) represent dimerization or higher oligomerization order (Hashimoto et al., 2011). In many cases, the biological activity of a protein complex is dependent on correct oligomeric order. Oligomerization may be required for a number of reasons, including the correct active site or cofactor binding site assembly on the oligomeric interface or allosteric regulation. Examples where dimerization is crucial for the formation of active sites on the oligomeric interface include previously mentioned aspartate aminotransferase (AspAT) [22], aspartate transcarbamoylase (ATC), and orotate phosphoribosyl transferase (PfOPRT) [183] from P. falciparum. In addition, the physiological assembly of PfOPRT/PfOPDC heterotetramer was shown to be more effective compared to the monofunctional enzymes [189]. A number of recent publications also suggest that the protein oligomerization to be a key driving force in evolution [190194].

Another important aspect of oligomerization is remarkable selectivity and binding affinity. Large surface area of the intraoligomeric interfaces and evolutional diversity allow oligomeric partners selectively bind to each other with no cross-reactivity in the system. In majority of cases, purification of oligomeric proteins from both native and recombinant sources can be performed without any foreign protein incorporations in the assembly. Unlike the active sites and cofactor binding sites where evolutionary constraints restrict the sequence diversity to retain the function, oligomeric interfaces are significantly less conserved among homologous proteins [195, 196]. Thus, small molecule compounds reacting with the conserved active site of target enzyme of the parasite will likely interact with the host’s homologous enzyme.

Direct interference with protein self-assembly would provide an opportunity for a highly selective modulation of protein activity or function both in vitro and in vivo.

Advertisement

9. Making (breaking) bad proteins

The recently proposed protein interference assay (PIA) [22] involves the utilization of structural knowledge (data) and mutagenic modification of one (or more) partner proteins in the assembly. These modifications may affect the binding site for a cofactor, catalytic activity, or disrupt the oligomeric interface of the target protein. Thus, recombinant and, most importantly, controlled co-expression of both wild type and its inactive (hyperactive) mutant would allow the formation of the complex with modified activity in vitro.

Previously mentioned homodimeric PfOPRT, as part of the PfOPRT/PfORDC heterotetramer, could also be a subject to PIA. The active sites of PfOPRT were reported to contain the amino acids from both subunits, suggesting that introduction of the active site mutants with modified activity in vivo would also affect the native PfOPRT. This assay would potentially bypass previously observed difficulties with poor inhibitor accessibility and aid in validation of the enzyme as antimalarial drug target.

Despite the obvious limitation of PIA approach to oligomeric proteins, this assay would still allow partial assessment of the system of interest, as many of the studied pathways are likely to involve at least one oligomeric assembly. We suggest that PIA would also allow re-evaluation of the previously studied promising targets where conventional validation approaches have failed.

Advertisement

10. Conclusion

In order to assess a gene’s product role, one must possess a set of tools, such as genetic manipulations (e.g. knockout, silencing etc.), to modulate the target function in vivo. Sufficient specificity (with little or no cross-reactivity) is essential for correct interpretation of the data. Although genetic manipulations have been proven to be highly effective in model and fully defined systems, less studied and complex systems remain highly challenging. In many pathogenic systems, including human malaria, conventional genetic manipulation techniques or small molecule inhibitor approaches do not always provide the desired efficacy [22]. In a number of human pathogens, multiple life cycle stages in different hosts and vectors make both in vitro and in vivo target characterization challenging to approach. A number of classic techniques such as silencing RNA [197, 198] have already been reported to be non-effective in certain cases [199202].

In addition, the use of small molecule inhibitor approaches in vivo is associated with high costs and is often limited due to the variety of host-specific reasons that are difficult to predict, such as rapid metabolism, poor membrane transport, or localization. For example, while a number of compounds were reported to inhibit PfOPRT activity in vitro as well as clear parasitemia in P. berghei-infected mice, in vivo trials with P. falciparum have failed [180]. Thus, potential drug targets may remain unexplored due to the inability to use the existing validation tool set.

Insufficient amount of effective target validation tools significantly limits the understanding of human pathogenic systems and hinders the rate of novel drug development. A constant supply of robust and effective techniques is needed in order to successfully dissect yet unexplored parasitic pathways, provide the basis for rational drug design, and counter-balance the ability of many human pathogens to rapidly develop drug resistance. We believe that protein interference assay (PIA) will enrich the currently available research toolset.

References

  1. 1. Whitfield J. Portrait of a serial killer. Nature News (2002), http://www.nature.com/news/2002/021003/full/news021001-6.html.
  2. 2. WHO. World Malaria Report. World Health Organisation. 2015 (Geneva, Switzerland).
  3. 3. Verlinden BK, Louw A, Birkholtz LM. Resisting resistance: is there a solution for malaria? Expert Opin Drug Discov. 2016;11(4):395-406.
  4. 4. Muller IB, Hyde JE. Antimalarial drugs: modes of action and mechanisms of parasite resistance. Future Microbiol. 2010;5(12):1857-1873.
  5. 5. Rodrigues T, Lopes F, Moreira R. Inhibitors of the mitochondrial electron transport chain and de novo pyrimidine biosynthesis as antimalarials: The present status. Curr Med Chem. 2010;17(10):929-956.
  6. 6. Biamonte MA, Wanner J, Le Roch KG. Recent advances in malaria drug discovery. Bioorg Med Chem Lett. 2013;23(10):2829-2843.
  7. 7. Cui L, Mharakurwa S, Ndiaye D, Rathod PK, Rosenthal PJ. Antimalarial drug resistance: literature review and activities and findings of the ICEMR Network. Am J Trop Med Hyg (2015). Vol 93(3 Suppl): 57–68.
  8. 8. Mbengue A, Bhattacharjee S, Pandharkar T, Liu H, Estiu G, Stahelin RV, et al. A molecular mechanism of artemisinin resistance in Plasmodium falciparum malaria. Nature. 2015;520(7549):683-687.
  9. 9. Paloque L, Ramadani AP, Mercereau-Puijalon O, Augereau JM, Benoit-Vical F. Plasmodium falciparum: multifaceted resistance to artemisinins. Malar J. 2016;15(1):149.
  10. 10. Avitia-Domínguez C, Sierra-Campos E, Betancourt-Conde I, Aguirre-Raudry M, Vázquez-Raygoza A, Luevano-De la Cruz A, et al. Targeting plasmodium metabolism to improve antimalarial drug design. Curr Protein Pept Sci. 2016;17(3):260-274.
  11. 11. Wells TN, Hooft van Huijsduijnen R, VanVoorhis WC. Malaria medicines: a glass half full? Nat Rev Drug Discov. 2015;14(6):424-442.
  12. 12. Isserlin R, Bader GD, Edwards A, Frye S, Willson T, Yu FH. The human genome and drug discovery after a decade. Roads (still) not taken. arXiv (2011) arXiv:1102.0448 [q-bio.OT].
  13. 13. Fitch CD. Ferriprotoporphyrin IX, phospholipids, and the antimalarial actions of quinoline drugs. Life Sci. 2004;74(16):1957-1972.
  14. 14. Cui L, Su XZ. Discovery, mechanisms of action and combination therapy of artemisinin. Expert Rev Anti Infect Ther. 2009;7(8):999-1013.
  15. 15. Carrington HC, Crowther AF, Davey DG, Levi AA, Rose FL. A metabolite of paludrine with high antimalarial activity. Nature. 1951;168(4288):1080.
  16. 16. Crowther AF, Levi AA. Proguanil, the isolation of a metabolite with high antimalarial activity. Br J Pharmacol Chemother. 1953;8(1):93-97.
  17. 17. Muregi FW. Antimalarial drugs and their useful therapeutic lives: rational drug design lessons from pleiotropic action of quinolines and artemisinins. Curr Drug Discov Technol. 2010;7(4):280-316.
  18. 18. Grueneberg DA, Degot S, Pearlberg J, Li W, Davies JE, Baldwin A, et al. Kinase requirements in human cells: I. Comparing kinase requirements across various cell types. Proc Natl Acad Sci U S A. 2008;105(43):16472-16477.
  19. 19. Fedorov O, Müller S, Knapp S. The (un)targeted cancer kinome. Nat Chem Biol. 2010;6(3):166-169.
  20. 20. Manning G, Whyte DB, Martinez R, Hunter T, Sudarsanam S. The protein kinase complement of the human genome. Science. 2002;298(5600):1912-1934.
  21. 21. Edwards AM, Isserlin R, Bader GD, Frye SV, Willson TM, Yu FH. Too many roads not taken. Nature. 2011;470(7333):163-165.
  22. 22. Meissner KA, Lunev S, Wang YZ, Linzke M, de Assis Batista F, Wrenger C, et al. Drug target validation methods in malaria—Protein interference assay (PIA) as a tool for highly specific drug target validation. Curr Drug Targets (2016). (in press).
  23. 23. Fry M, Pudney M. Site of action of the antimalarial hydroxynaphthoquinone, 2-[trans-4-(4'-chlorophenyl) cyclohexyl]-3-hydroxy-1,4-naphthoquinone (566C80). Biochemical Pharmacology. 1992;43(7):1545-1553.
  24. 24. Baldwin J, Farajallah AM, Malmquist NA, Rathod PK, Phillips MA. Malarial dihydroorotate dehydrogenase. Substrate and inhibitor specificity. J Biol Chem. 2002;277(44):41827-41834.
  25. 25. Baldwin J, Michnoff CH, Malmquist NA, White J, Roth MG, Rathod PK, et al. High-throughput screening for potent and selective inhibitors of Plasmodium falciparum dihydroorotate dehydrogenase. J Biol Chem. 2005;280(23):21847-21853.
  26. 26. Phillips MA, Lotharius J, Marsh K, White J, Dayan A, White KL, et al. A long-duration dihydroorotate dehydrogenase inhibitor (DSM265) for prevention and treatment of malaria. Sci Transl Med. 2015;7(296):296ra111.
  27. 27. Pavadai E, El Mazouni F, Wittlin S, de Kock C, Phillips MA, Chibale K. Identification of new human malaria parasite Plasmodium falciparum dihydroorotate dehydrogenase inhibitors by pharmacophore and structure-based virtual screening. J Chem Inf Model. (2016). Vol 56(3):548-62.
  28. 28. Robert A, Benoit-Vical F, Claparols C, Meunier B. The antimalarial drug artemisinin alkylates heme in infected mice. Proc Natl Acad Sci U S A. 2005;102(38):13676-13680.
  29. 29. Nosten F, White NJ. Artemisinin-based combination treatment of falciparum malaria. Am J Trop Med Hyg. 2007;77(6 Suppl):181-192.
  30. 30. Ashley EA, Dhorda M, Fairhurst RM, Amaratunga C, Lim P, Suon S, et al. Spread of artemisinin resistance in Plasmodium falciparum malaria. N Engl J Med. 2014;371(5):411-423.
  31. 31. White N. Antimalarial drug resistance and combination chemotherapy. Philos Trans R Soc Lond B Biol Sci. 1999;354(1384):739-749.
  32. 32. Ke H, Morrisey JM, Ganesan SM, Painter HJ, Mather MW, Vaidya AB. Variation among Plasmodium falciparum strains in their reliance on mitochondrial electron transport chain function. Eukaryotic cell. 2011;10(8):1053-1061.
  33. 33. Vaidya AB, Mather MW. Mitochondrial evolution and functions in malaria parasites. Annu Rev Microbiol 2009;63:249-267.
  34. 34. Ke H, Lewis IA, Morrisey JM, McLean KJ, Ganesan SM, Painter HJ, et al. Genetic investigation of tricarboxylic acid metabolism during the Plasmodium falciparum life cycle. Cell Rep. 2015;11(1):164-174.
  35. 35. MacRae JI, Dixon MW, Dearnley MK, Chua HH, Chambers JM, Kenny S, et al. Mitochondrial metabolism of sexual and asexual blood stages of the malaria parasite Plasmodium falciparum. BMC biology. 2013;11:67.
  36. 36. Bryant C, Voller A, Smith MJ. The incorporation of radioactivity from (14C) glucose into the soluble metabolic intermediates of malaria parasites. Am J Trop Med Hyg. 1964;13:515-519.
  37. 37. Scheibel LW, Pflaum WK. Cytochrome oxidase activity in platelet-free preparations of Plasmodium falciparum. J Parasitol. 1970;56(6):1054.
  38. 38. Roth EF, Calvin MC, Max-Audit I, Rosa J, Rosa R. The enzymes of the glycolytic pathway in erythrocytes infected with Plasmodium falciparum malaria parasites. Blood. 1988;72(6):1922-1925.
  39. 39. Roth EF, Raventos-Suarez C, Perkins M, Nagel RL. Glutathione stability and oxidative stress in P. falciparum infection in vitro: responses of normal and G6PD deficient cells. Biochem Biophys Res Commun. 1982;109(2):355-362.
  40. 40. Planche T, Krishna S. Severe malaria: metabolic complications. Curr Mol Med. 2006;6(2):141-153.
  41. 41. Fleck SL, Pudney M, Sinden RE. The effect of atovaquone (566C80) on the maturation and viability of Plasmodium falciparum gametocytes in vitro. Trans R Soc Trop Med Hyg. 1996;90(3):309-312.
  42. 42. Srivastava IK, Rottenberg H, Vaidya AB. Atovaquone, a broad spectrum antiparasitic drug, collapses mitochondrial membrane potential in a malarial parasite. J Biol Chem. 1997;272(7):3961-3966.
  43. 43. Mather MW, Darrouzet E, Valkova-Valchanova M, Cooley JW, McIntosh MT, Daldal F, et al. Uncovering the molecular mode of action of the antimalarial drug atovaquone using a bacterial system. J Biol Chem. 2005;280(29):27458-27465.
  44. 44. Nixon GL, Pidathala C, Shone AE, Antoine T, Fisher N, O'Neill PM, et al. Targeting the mitochondrial electron transport chain of Plasmodium falciparum: New strategies towards the development of improved antimalarials for the elimination era. Future Med Chem. 2013;5(13):1573-1591.
  45. 45. Fisher N, Bray PG, Ward SA, Biagini GA. The malaria parasite type II NADH:quinone oxidoreductase: an alternative enzyme for an alternative lifestyle. Trends Parasitol. 2007;23(7):305-310.
  46. 46. Fry M, Webb E, Pudney M. Effect of mitochondrial inhibitors on adenosinetriphosphate levels in Plasmodium falciparum. Comp Biochem Physiol B Comp Biochem. 1990;96(4):775-782.
  47. 47. Balabaskaran Nina P, Morrisey JM, Ganesan SM, Ke H, Pershing AM, Mather MW, et al. ATP synthase complex of Plasmodium falciparum: dimeric assembly in mitochondrial membranes and resistance to genetic disruption. J Biol Chem. 2011;286(48):41312-41322.
  48. 48. Phillips MA, Rathod PK. Plasmodium dihydroorotate dehydrogenase: a promising target for novel anti-malarial chemotherapy. Infect Disord Drug Targets. 2010;10(3):226-239.
  49. 49. Gutteridge WE, Dave D, Richards WH. Conversion of dihydroorotate to orotate in parasitic protozoa. Biochim et Biophys Acta. 1979;582(3):390-401.
  50. 50. Stocks PA, Barton V, Antoine T, Biagini GA, Ward SA, O'Neill PM. Novel inhibitors of the Plasmodium falciparum electron transport chain. Parasitology. 2014;141(1):50-65.
  51. 51. Herrmann ML, Schleyerbach R, Kirschbaum BJ. Leflunomide: an immunomodulatory drug for the treatment of rheumatoid arthritis and other autoimmune diseases. Immunopharmacology. 2000;47(2-3):273-289.
  52. 52. Shannon PVRE, Eichholtz T., Linstead D, Masdin P, Skinner R, inventor. Condensed heterocyclic compounds as anti-inflammatory and immunomodulatory agents, 1999. Google Patents, WO1999045926B1.
  53. 53. Boa AN, Canavan SP, Hirst PR, Ramsey C, Stead AM, McConkey GA. Synthesis of brequinar analogue inhibitors of malaria parasite dihydroorotate dehydrogenase. Bioorg Med Chem. 2005;13(6):1945-1967.
  54. 54. Patel V, Booker M, Kramer M, Ross L, Celatka CA, Kennedy LM, et al. Identification and characterization of small molecule inhibitors of Plasmodium falciparum dihydroorotate dehydrogenase. J Biol Chem. 2008;283(50):35078-35085.
  55. 55. Booker ML, Bastos CM, Kramer ML, Barker RH, Jr., Skerlj R, Sidhu AB, et al. Novel inhibitors of Plasmodium falciparum dihydroorotate dehydrogenase with anti-malarial activity in the mouse model. J Biol Chem. 2010;285(43):33054-33064.
  56. 56. Skerlj RT, Bastos CM, Booker ML, Kramer ML, Barker RH, Jr., Celatka CA, et al. Optimization of potent inhibitors of P. falciparum dihydroorotate dehydrogenase for the treatment of malaria. ACS Med Chem Lett. 2011;2(9):708-713.
  57. 57. Bedingfield PT, Cowen D, Acklam P, Cunningham F, Parsons MR, McConkey GA, et al. Factors influencing the specificity of inhibitor binding to the human and malaria parasite dihydroorotate dehydrogenases. J Med Chem. 2012;55(12):5841-5850.
  58. 58. Fritzson I BP, Sundin AP, McConkey G, Nilsson UJ. N-substituted salicylamides as selective malaria parasite dihydroorotate dehydrogenase inhibitors. Med Chem Commun. 2011;2:3.
  59. 59. Davies M, Heikkila T, McConkey GA, Fishwick CW, Parsons MR, Johnson AP. Structure-based design, synthesis, and characterization of inhibitors of human and Plasmodium falciparum dihydroorotate dehydrogenases. J Med Chem. 2009;52(9):2683-2693.
  60. 60. Phillips MA, Gujjar R, Malmquist NA, White J, El Mazouni F, Baldwin J, et al. Triazolopyrimidine-based dihydroorotate dehydrogenase inhibitors with potent and selective activity against the malaria parasite Plasmodium falciparum. J Med Chem. 2008;51(12):3649-3653.
  61. 61. Deng X, Gujjar R, El Mazouni F, Kaminsky W, Malmquist NA, Goldsmith EJ, et al. Structural plasticity of malaria dihydroorotate dehydrogenase allows selective binding of diverse chemical scaffolds. J Biol Chem. 2009;284(39):26999-27009.
  62. 62. Gujjar R, Marwaha A, El Mazouni F, White J, White KL, Creason S, et al. Identification of a metabolically stable triazolopyrimidine-based dihydroorotate dehydrogenase inhibitor with antimalarial activity in mice. J Med Chem. 2009;52(7):1864-1872.
  63. 63. Gujjar R, El Mazouni F, White KL, White J, Creason S, Shackleford DM, et al. Lead optimization of aryl and aralkyl amine-based triazolopyrimidine inhibitors of Plasmodium falciparum dihydroorotate dehydrogenase with antimalarial activity in mice. J Med Chem. 2011;54(11):3935-3949.
  64. 64. Marwaha A, White J, El Mazouni F, Creason SA, Kokkonda S, Buckner FS, et al. Bioisosteric transformations and permutations in the triazolopyrimidine scaffold to identify the minimum pharmacophore required for inhibitory activity against Plasmodium falciparum dihydroorotate dehydrogenase. J Med Chem. 2012;55(17):7425-7436.
  65. 65. Schutz M, Brugna M, Lebrun E, Baymann F, Huber R, Stetter KO, et al. Early evolution of cytochrome bc complexes. J Mol Biol. 2000;300(4):663-675.
  66. 66. Gao X, Wen X, Yu C, Esser L, Tsao S, Quinn B, et al. The crystal structure of mitochondrial cytochrome bc1 in complex with famoxadone: the role of aromatic-aromatic interaction in inhibition. Biochemistry. 2002;41(39):11692-11702.
  67. 67. Berry EA, Guergova-Kuras M, Huang LS, Crofts AR. Structure and function of cytochrome bc complexes. Annu Rev Biochem. 2000;69:1005-1075.
  68. 68. Iwata S, Lee JW, Okada K, Lee JK, Iwata M, Rasmussen B, et al. Complete structure of the 11-subunit bovine mitochondrial cytochrome bc1 complex. Science. 1998;281(5373):64-71.
  69. 69. Mitchell P. Possible molecular mechanisms of the protonmotive function of cytochrome systems. J Theor Biol. 1976;62(2):327-367.
  70. 70. Rieske JS, Zaugg WS, Hansen RE. Studies on the electron transfer system. Lix. Distribution of iron and of the component giving an electron paramagnetic resonance signal at G = 1.90 in subfractions of complex 3. J Biol Chem. 1964;239:3023-3030.
  71. 71. Berry EA, Huang LS. Conformationally linked interaction in the cytochrome bc[1] complex between inhibitors of the Q(o) site and the Rieske iron-sulfur protein. Biochim Biophys Acta. 2011;1807(10):1349-1363.
  72. 72. Vaidya AB. Mitochondrial and plastid functions as antimalarial drug targets. Curr Drug Targets Infect Disord. 2004;4(1):11-23.
  73. 73. Hunte C, Koepke J, Lange C, Rossmanith T, Michel H. Structure at 2.3 A resolution of the cytochrome bc(1) complex from the yeast Saccharomyces cerevisiae co-crystallized with an antibody Fv fragment. Structure. 2000;8(6):669-684.
  74. 74. Srivastava IK, Vaidya AB. A mechanism for the synergistic antimalarial action of atovaquone and proguanil. Antimicrob Agents Chemother. 1999;43(6):1334-1339.
  75. 75. Birth D, Kao WC, Hunte C. Structural analysis of atovaquone-inhibited cytochrome bc1 complex reveals the molecular basis of antimalarial drug action. Nature Commun 2014;5:4029.
  76. 76. Looareesuwan S, Viravan C, Webster HK, Kyle DE, Hutchinson DB, Canfield CJ. Clinical studies of atovaquone, alone or in combination with other antimalarial drugs, for treatment of acute uncomplicated malaria in Thailand. Am J Trop Med Hyg. 1996;54(1):62-66.
  77. 77. Brunton LL, Chabner BA, Knollman BC. Goodman and Gilman’s The Pharmacological Basis of Therapeutics. 11th ed. New York: McGraw Hill; 2011.
  78. 78. Winter RW, Kelly JX, Smilkstein MJ, Dodean R, Bagby GC, Rathbun RK, et al. Evaluation and lead optimization of anti-malarial acridones. Exp Parasitol. 2006;114(1):47-56.
  79. 79. Nilsen A, Miley GP, Forquer IP, Mather MW, Katneni K, Li Y, et al. Discovery, synthesis, and optimization of antimalarial 4(1H)-quinolone-3-diarylethers. J Med Chem. 2014;57(9):3818-3834.
  80. 80. Nilsen A, LaCrue AN, White KL, Forquer IP, Cross RM, Marfurt J, et al. Quinolone-3-diarylethers: a new class of antimalarial drug. Sci Transl Med. 2013;5(177):177ra37.
  81. 81. Biagini GA, Fisher N, Shone AE, Mubaraki MA, Srivastava A, Hill A, et al. Generation of quinolone antimalarials targeting the Plasmodium falciparum mitochondrial respiratory chain for the treatment and prophylaxis of malaria. Proc Natl Acad Sci USA. 2012;109(21):8298-8303.
  82. 82. Yeates CL, Batchelor JF, Capon EC, Cheesman NJ, Fry M, Hudson AT, et al. Synthesis and structure-activity relationships of 4-pyridones as potential antimalarials. J Med Chem. 2008;51(9):2845-2852.
  83. 83. Bueno JM, Herreros E, Angulo-Barturen I, Ferrer S, Fiandor JM, Gamo FJ, et al. Exploration of 4(1H)-pyridones as a novel family of potent antimalarial inhibitors of the plasmodial cytochrome bc1. Future Med Chem. 2012;4(18):2311-2323.
  84. 84. Lukens AK, Heidebrecht RW, Jr., Mulrooney C, Beaudoin JA, Comer E, Duvall JR, et al. Diversity-oriented synthesis probe targets Plasmodium falciparum cytochrome b ubiquinone reduction site and synergizes with oxidation site inhibitors. J Infect Dis. 2015;211(7):1097-1103.
  85. 85. Esser L, Quinn B, Li YF, Zhang M, Elberry M, Yu L, et al. Crystallographic studies of quinol oxidation site inhibitors: a modified classification of inhibitors for the cytochrome bc(1) complex. J Mol Biol. 2004;341(1):281-302.
  86. 86. Gao X, Wen X, Esser L, Quinn B, Yu L, Yu CA, et al. Structural basis for the quinone reduction in the bc1 complex: a comparative analysis of crystal structures of mitochondrial cytochrome bc1 with bound substrate and inhibitors at the Qi site. Biochemistry. 2003;42(30):9067-9080.
  87. 87. Li H, Zhu XL, Yang WC, Yang GF. Comparative kinetics of Qi site inhibitors of cytochrome bc1 complex: picomolar antimycin and micromolar cyazofamid. Chem Biol Drug Design. 2014;83(1):71-80.
  88. 88. Berry EA, Huang LS, Lee DW, Daldal F, Nagai K, Minagawa N. Ascochlorin is a novel, specific inhibitor of the mitochondrial cytochrome bc1 complex. Biochim Biophys Acta. 2010;1797(3):360-370.
  89. 89. Capper MJ, O'Neill PM, Fisher N, Strange RW, Moss D, Ward SA, et al. Antimalarial 4(1H)-pyridones bind to the Qi site of cytochrome bc1. Proc Natl Acad Sci USA. 2015;112(3):755-760.
  90. 90. Biagini GA, Viriyavejakul P, O'neill PM, Bray PG, Ward SA. Functional characterization and target validation of alternative complex I of Plasmodium falciparum mitochondria. Antimicrob Agents Chemother. 2006;50(5):1841-1851.
  91. 91. Kerscher SJ. Diversity and origin of alternative NADH:ubiquinone oxidoreductases. Biochim Biophys Acta. 2000;1459(2-3):274-283.
  92. 92. Luttik MA, Overkamp KM, Kotter P, de Vries S, van Dijken JP, Pronk JT. The Saccharomyces cerevisiae NDE1 and NDE2 genes encode separate mitochondrial NADH dehydrogenases catalyzing the oxidation of cytosolic NADH. J Biol Chem. 1998;273(38):24529-24534.
  93. 93. Marres CA, de Vries S, Grivell LA. Isolation and inactivation of the nuclear gene encoding the rotenone-insensitive internal NADH: ubiquinone oxidoreductase of mitochondria from Saccharomyces cerevisiae. Eur J Biochem/FEBS. 1991;195(3):857-862.
  94. 94. Melo AM, Bandeiras TM, Teixeira M. New insights into type II NAD(P)H:quinone oxidoreductases. Microbiol Mol Biol Rev: MMBR. 2004;68(4):603-616.
  95. 95. Rasmusson AG, Soole KL, Elthon TE. Alternative NAD(P)H dehydrogenases of plant mitochondria. Annu Rev Plant Biol. 2004;55:23-39.
  96. 96. Yagi T. Bacterial NADH-quinone oxidoreductases. J Bioenerg Biomembr. 1991;23(2):211-225.
  97. 97. Mattevi A, Obmolova G, Sokatch JR, Betzel C, Hol WG. The refined crystal structure of Pseudomonas putida lipoamide dehydrogenase complexed with NAD+ at 2.45 A resolution. Proteins. 1992;13(4):336-351.
  98. 98. Boysen KE, Matuschewski K. Arrested oocyst maturation in Plasmodium parasites lacking type II NADH:ubiquinone dehydrogenase. J Biol Chem. 2011;286(37):32661-32671.
  99. 99. Yeh I, Hanekamp T, Tsoka S, Karp PD, Altman RB. Computational analysis of Plasmodium falciparum metabolism: organizing genomic information to facilitate drug discovery. Genome Res. 2004;14(5):917-924.
  100. 100. Dong CK, Patel V, Yang JC, Dvorin JD, Duraisingh MT, Clardy J, et al. Type II NADH dehydrogenase of the respiratory chain of Plasmodium falciparum and its inhibitors. Bioorg Med Chem Lett. 2009;19(3):972-975.
  101. 101. Eschemann A, Galkin A, Oettmeier W, Brandt U, Kerscher S. HDQ (1-hydroxy-2-dodecyl-4(1H)quinolone), a high affinity inhibitor for mitochondrial alternative NADH dehydrogenase: evidence for a ping-pong mechanism. J Biol Chem. 2005;280(5):3138-3142.
  102. 102. Saleh A, Friesen J, Baumeister S, Gross U, Bohne W. Growth inhibition of Toxoplasma gondii and Plasmodium falciparum by nanomolar concentrations of 1-hydroxy-2-dodecyl-4(1H)quinolone, a high-affinity inhibitor of alternative (type II) NADH dehydrogenases. Antimicrob Agents Chemother. 2007;51(4):1217-1222.
  103. 103. Vallieres C, Fisher N, Antoine T, Al-Helal M, Stocks P, Berry NG, et al. HDQ, a potent inhibitor of Plasmodium falciparum proliferation, binds to the quinone reduction site of the cytochrome bc1 complex. Antimicrob Agents Chemother. 2012;56(7):3739-3747.
  104. 104. Li W, Mo W, Shen D, Sun L, Wang J, Lu S, et al. Yeast model uncovers dual roles of mitochondria in action of artemisinin. PLoS Genet. 2005;1(3):e36.
  105. 105. Antoine T, Fisher N, Amewu R, O'Neill PM, Ward SA, Biagini GA. Rapid kill of malaria parasites by artemisinin and semi-synthetic endoperoxides involves ROS-dependent depolarization of the membrane potential. J Antimicrob Chemother. 2014;69(4):1005-1016.
  106. 106. Pidathala C, Amewu R, Pacorel B, Nixon GL, Gibbons P, Hong WD, et al. Identification, design and biological evaluation of bisaryl quinolones targeting Plasmodium falciparum type II NADH:quinone oxidoreductase (PfNDH2). J Med Chem. 2012;55(5):1831-1843.
  107. 107. Leung SC, Gibbons P, Amewu R, Nixon GL, Pidathala C, Hong WD, et al. Identification, design and biological evaluation of heterocyclic quinolones targeting Plasmodium falciparum type II NADH:quinone oxidoreductase (PfNDH2). J Med Chem. 2012;55(5):1844-1857.
  108. 108. Klingenberg M. Localization of the glycerol-phosphate dehydrogenase in the outer phase of the mitochondrial inner membrane. Eur J Biochem/FEBS. 1970;13(2):247-252.
  109. 109. Yeh JI, Chinte U, Du S. Structure of glycerol-3-phosphate dehydrogenase, an essential monotopic membrane enzyme involved in respiration and metabolism. Proc Natl Acad Sci USA. 2008;105(9):3280-3285.
  110. 110. Fry M, Beesley JE. Mitochondria of mammalian Plasmodium spp. Parasitology. 1991;102 Pt 1:17-26.
  111. 111. Uyemura SA, Luo S, Moreno SN, Docampo R. Oxidative phosphorylation, Ca(2+) transport, and fatty acid-induced uncoupling in malaria parasites mitochondria. J Biol Chem. 2000;275(13):9709-9715.
  112. 112. Uyemura SA, Luo S, Vieira M, Moreno SN, Docampo R. Oxidative phosphorylation and rotenone-insensitive malate- and NADH-quinone oxidoreductases in Plasmodium yoelii yoelii mitochondria in situ. J Biol Chem. 2004;279(1):385-393.
  113. 113. Hatefi Y. The mitochondrial electron transport and oxidative phosphorylation system. Annu Rev Biochem. 1985;54:1015-1069.
  114. 114. Kita K, Vibat CR, Meinhardt S, Guest JR, Gennis RB. One-step purification from Escherichia coli of complex II (succinate: ubiquinone oxidoreductase) associated with succinate-reducible cytochrome b556. J Biol Chem. 1989;264(5):2672-2677.
  115. 115. Moll R, Schafer G. Purification and characterisation of an archaebacterial succinate dehydrogenase complex from the plasma membrane of the thermoacidophile Sulfolobus acidocaldarius. Eur J Biochem/FEBS. 1991;201(3):593-600.
  116. 116. Reddy TL, Weber MM. Solubilization, purification, and characterization of succinate dehydrogenase from membranes of Mycobacterium phlei. J Bacteriol. 1986;167(1):1-6.
  117. 117. McNeil MB, Hampton HG, Hards KJ, Watson BN, Cook GM, Fineran PC. The succinate dehydrogenase assembly factor, SdhE, is required for the flavinylation and activation of fumarate reductase in bacteria. FEBS Lett. 2014;588(3):414-421.
  118. 118. Hartman T, Weinrick B, Vilcheze C, Berney M, Tufariello J, Cook GM, et al. Succinate dehydrogenase is the regulator of respiration in Mycobacterium tuberculosis. PLoS Pathog. 2014;10(11):e1004510.
  119. 119. Kita K, Oya H, Gennis RB, Ackrell BA, Kasahara M. Human complex II (succinate-ubiquinone oxidoreductase): cDNA cloning of iron sulfur (Ip) subunit of liver mitochondria. Biochem Biophys Res Commun. 1990;166(1):101-108.
  120. 120. Hatefi Y, Stiggall DL. Preparation and properties of succinate: ubiquinone oxidoreductase (complex II). Methods Enzymol. 1978;53:21-27.
  121. 121. Tushurashvili PR, Gavrikova EV, Ledenev AN, Vinogradov AD. Studies on the succinate dehydrogenating system. Isolation and properties of the mitochondrial succinate-ubiquinone reductase. Biochim Biophys Acta. 1985;809(2):145-159.
  122. 122. Suraveratum N, Krungkrai SR, Leangaramgul P, Prapunwattana P, Krungkrai J. Purification and characterization of Plasmodium falciparum succinate dehydrogenase. Mol Biochem Parasitol. 2000;105(2):215-222.
  123. 123. Takeo S, Kokaze A, Ng CS, Mizuchi D, Watanabe JI, Tanabe K, et al. Succinate dehydrogenase in Plasmodium falciparum mitochondria: molecular characterization of the SDHA and SDHB genes for the catalytic subunits, the flavoprotein (Fp) and iron-sulfur (Ip) subunits. Mol Biochem Parasitol. 2000;107(2):191-205.
  124. 124. Hagerhall C. Succinate: quinone oxidoreductases. Variations on a conserved theme. Biochim Biophys Acta. 1997;1320(2):107-141.
  125. 125. Kita K, Hirawake H, Miyadera H, Amino H, Takeo S. Role of complex II in anaerobic respiration of the parasite mitochondria from Ascaris suum and Plasmodium falciparum. Biochim Biophys Acta. 2002;1553(1-2):123-139.
  126. 126. Molenaar D, van der Rest ME, Petrovic S. Biochemical and genetic characterization of the membrane-associated malate dehydrogenase (acceptor) from Corynebacterium glutamicum. Eur J Biochem/FEBS. 1998;254(2):395-403.
  127. 127. Mogi T, Murase Y, Mori M, Shiomi K, Omura S, Paranagama MP, et al. Polymyxin B identified as an inhibitor of alternative NADH dehydrogenase and malate: quinone oxidoreductase from the Gram-positive bacterium Mycobacterium smegmatis. J Biochem. 2009;146(4):491-499.
  128. 128. van Dooren GG, Marti M, Tonkin CJ, Stimmler LM, Cowman AF, McFadden GI. Development of the endoplasmic reticulum, mitochondrion and apicoplast during the asexual life cycle of Plasmodium falciparum. Mol Microbiol. 2005;57(2):405-419.
  129. 129. Guillemont J, Meyer C, Poncelet A, Bourdrez X, Andries K. Diarylquinolines, synthesis pathways and quantitative structure--activity relationship studies leading to the discovery of TMC207. Future Med Chem. 2011;3(11):1345-1360.
  130. 130. Basco LK, Le Bras J. In vitro activity of mitochondrial ATP synthetase inhibitors against Plasmodium falciparum. J Eukaryot Microbiol. 1994;41(3):179-183.
  131. 131. Sturm A, Mollard V, Cozijnsen A, Goodman CD, McFadden GI. Mitochondrial ATP synthase is dispensable in blood-stage Plasmodium berghei rodent malaria but essential in the mosquito phase. Proc Natl Acad Sci USA. 2015;112(33):10216-10223.
  132. 132. Olszewski KL, Mather MW, Morrisey JM, Garcia BA, Vaidya AB, Rabinowitz JD, et al. Branched tricarboxylic acid metabolism in Plasmodium falciparum. Nature. 2010;466(7307):774-778.
  133. 133. Olszewski KL, Mather MW, Morrisey JM, Garcia BA, Vaidya AB, Rabinowitz JD, et al. Retraction: Branched tricarboxylic acid metabolism in Plasmodium falciparum. Nature. 2013;497(7451):652.
  134. 134. Olszewski KL, Llinas M. Central carbon metabolism of Plasmodium parasites. Mol Biochem Parasitol. 2011;175(2):95-103.
  135. 135. Saunders EC, Ng WW, Chamber JM, Ng M, Naderer T, Kroemer JO, et al. Isoptopomer profiling of Leishmania mexicana promastigotes reveals important roles for succinate fermentation and aspartate uptake in TCA cycle anaplerosis, glutamate synthesis and growth. J Biol Chem. 2011;286.
  136. 136. Jäger J, Moser M, Sauder U, Jansonius JN. Crystal structures of Escherichia coli aspartate aminotransferase in two conformations. Comparison of an unliganded open and two liganded closed forms. J Mol Biol. 1994;239(2):285-305.
  137. 137. Kamitori S, Okamoto A, Hirotsu K, Higuchi T, Kuramitsu S, Kagamiyama H, et al. Three-dimensional structures of aspartate aminotransferase from Escherichia coli and its mutant enzyme at 2.5 A resolution. J Biochem. 1990;108(2):175-184.
  138. 138. Okamoto A, Higuchi T, Hirotsu K, Kuramitsu S, Kagamiyama H. X-ray crystallographic study of pyridoxal 5'-phosphate-type aspartate aminotransferases from Escherichia coli in open and closed form. J Biochem. 1994;116(1):95-107.
  139. 139. Smith DL, Almo SC, Toney MD, Ringe D. 2.8-A-resolution crystal structure of an active-site mutant of aspartate aminotransferase from Escherichia coli. Biochemistry. 1989;28(20):8161-8167.
  140. 140. Jeffrey PD, Bewley MC, MacGillivray RT, Mason AB, Woodworth RC, Baker EN. Ligand-induced conformational change in transferrins: crystal structure of the open form of the N-terminal half-molecule of human transferrin. Biochemistry. 1998;37(40):13978-13986.
  141. 141. Arnone A, Rogers PH, Hyde CC, Briley PD, Metzler CM, Metzler DE. In Transaminases (eds P. Christen and D. Metzler), pp. 138-155 John Wiley and Sons, NY. 1985.
  142. 142. Ford GC, Eichele G, Jansonius JN. Three-dimensional structure of a pyridoxal-phosphate-dependent enzyme, mitochondrial aspartate aminotransferase. Proc Natl Acad Sci U S A. 1980;77(5):2559-2563.
  143. 143. Malashkevich VN, Toney MD, Jansonius JN. Crystal structures of true enzymatic reaction intermediates: aspartate and glutamate ketimines in aspartate aminotransferase. Biochemistry. 1993;32(49):13451-13462.
  144. 144. McPhalen CA, Vincent MG, Jansonius JN. X-ray structure refinement and comparison of three forms of mitochondrial aspartate aminotransferase. J Mol Biol. 1992;225(2):495-517.
  145. 145. Jain R, Jordanova R, Muller IB, Wrenger C, Groves MR. Purification, crystallization and preliminary X-ray analysis of the aspartate aminotransferase of Plasmodium falciparum. Acta Crystallogr Sect F Struct Biol Cryst Commun 2010;66(Pt 4):409-412.
  146. 146. Wrenger C, Muller IB, Schifferdecker AJ, Jain R, Jordanova R, Groves MR. Specific inhibition of the aspartate aminotransferase of Plasmodium falciparum. J Mol Bio wl. 2011;405(4):956-971.
  147. 147. McPhalen CA, Vincent MG, Picot D, Jansonius JN, Lesk AM, Chothia C. Domain closure in mitochondrial aspartate aminotransferase. J Mol Biol. 1992;227(1):197-213.
  148. 148. Battchikova N, Koivulehto M, Denesyuk A, Ptitsyn L, Boretsky Y, Hellman J, et al. Aspartate aminotransferase from an alkalophilic Bacillus contains an additional 20-amino acid extension at its functionally important N-terminus. J Biochem. 1996;120(2): 425-432.
  149. 149. Wrenger C, Müller IB, Butzloff S, Jordanova R, Lunev S, Groves MR. Crystallization and preliminary X-ray diffraction of malate dehydrogenase from Plasmodium falciparum. Acta Crystallogr Sect F Struct Biol Cryst Commun. 2012;68(Pt 6):659-662.
  150. 150. Lunev S, Groves MR, Müller IB, Butzloff S, Wrenger C. Oligomeric protein interference demonstrates a measurable phenotype in cultured malarial parasites. Manuscript in preparation.
  151. 151. Woods SA, Schwartzbach SD, Guest JR. Two biochemically distinct classes of fumarase in Escherichia coli. Biochim Biophy Acta. 1988;954(1):14-26.
  152. 152. Flint DH, Emptage MH, Guest JR. Fumarase a from Escherichia coli: Purification and characterization as an iron-sulfur cluster containing enzyme. Biochemistry. 1992;31(42):10331-10337.
  153. 153. Bulusu V, Jayaraman V, Balaram H. Metabolic fate of fumarate, a side product of the purine salvage pathway in the intraerythrocytic stages of Plasmodium falciparum. J Biol Chem. 2011;286(11):9236-9245.
  154. 154. White NJ, Pukrittayakamee S, Hien TT, Faiz MA, Mokuolu OA, Dondorp AM. Malaria. Lancet. 2014;383(9918):723-735.
  155. 155. Hyde JE. Targeting purine and pyrimidine metabolism in human apicomplexan parasites. Curr Drug Targets. 2007;8(1):31-47.
  156. 156. Loffler M, Fairbanks LD, Zameitat E, Marinaki AM, Simmonds HA. Pyrimidine pathways in health and disease. Trends Mol Med. 2005;11(9):430-437.
  157. 157. Reyes P, Rathod PK, Sanchez DJ, Mrema JE, Rieckmann KH, Heidrich HG. Enzymes of purine and pyrimidine metabolism from the human malaria parasite, Plasmodium falciparum. Mol Biochem Parasitol. 1982;5(5):275-290.
  158. 158. Gardner MJ, Hall N, Fung E, White O, Berriman M, Hyman RW, et al. Genome sequence of the human malaria parasite Plasmodium falciparum. Nature. 2002;419(6906):498-511.
  159. 159. Cassera MB, Zhang Y, Hazleton KZ, Schramm VL. Purine and pyrimidine pathways as targets in Plasmodium falciparum. Curr Top Med Chem. 2011;11(16):2103-2115.
  160. 160. Gero AM, Brown GV, O'Sullivan WJ. Pyrimidine de novo synthesis during the life cycle of the intraerythrocytic stage of Plasmodium falciparum. J Parasitol. 1984;70(4):536-541.
  161. 161. Flores MV, O'Sullivan WJ, Stewart TS. Characterisation of the carbamoyl phosphate synthetase gene from Plasmodium falciparum. Mol Biochem Parasitol. 1994;68(2):315-318.
  162. 162. Flores MV, Atkins D, Wade D, O'Sullivan WJ, Stewart TS. Inhibition of Plasmodium falciparum proliferation in vitro by ribozymes. J Biol Chem. 1997;272(27):16940-16945.
  163. 163. Hendry P, McCall MJ, Stewart TS, Lockett TJ. Redesigned and chemically-modified hammerhead ribozymes with improved activity and serum stability. BMC chemical biology. 2004;4(1):1.
  164. 164. Alvarez-Salas LM. Nucleic acids as therapeutic agents. Curr Topics Med Chem. 2008;8(15):1379-1404.
  165. 165. Madani S, Baillon J, Fries J, Belhadj O, Bettaieb A, Ben Hamida M, et al. Pyrimidine pathways enzymes in human tumors of brain and associated tissues: Potentialities for the therapeutic use of N-(phosphonacetyl-l-aspartate and 1-beta-d-arabinofuranosylcytosine. Eur J Cancer Clin Oncol. 1987;23(10):1485-1490.
  166. 166. Banerjee A, Arora N, Murty U. Aspartate carbamoyltransferase of Plasmodium falciparum as a potential drug target for designing anti-malarial chemotherapeutic agents. Med Chem Res. 2012;21(9):2480-2493.
  167. 167. Depamede SN, Menz I. Phylogenetic analysis and protein modeling of Plasmodium falciparum aspartate transcarbamoylase (ATCase). Res J Microbiol. 2011;6:599-608.
  168. 168. Sun W, Tanaka TQ, Magle CT, Huang W, Southall N, Huang R, et al. Chemical signatures and new drug targets for gametocytocidal drug development. Sci Rep. 2014;4:3743.
  169. 169. Lunev S, Bosch SS, Batista FdeA, Wrenger C, Groves MR. Crystal structure of truncated aspartate transcarbamoylase from Plasmodium falciparum. Acta Crystallogr F Struct Biol Commun. 2016;72(Pt 7):523-533.
  170. 170. Simmer JP, Kelly RE, Rinker AG, Jr., Zimmermann BH, Scully JL, Kim H, et al. Mammalian dihydroorotase: nucleotide sequence, peptide sequences, and evolution of the dihydroorotase domain of the multifunctional protein CAD. Proc Natl Acad Sci USA. 1990;87(1):174-178.
  171. 171. Krungkrai J, Krungkrai SR, Phakanont K. Antimalarial activity of orotate analogs that inhibit dihydroorotase and dihydroorotate dehydrogenase. Biochem Pharmacol. 1992;43(6):1295-1301.
  172. 172. Seymour KK, Lyons SD, Phillips L, Rieckmann KH, Christopherson RI. Cytotoxic effects of inhibitors of de novo pyrimidine biosynthesis upon Plasmodium falciparum. Biochemistry. 1994;33(17):5268-5274.
  173. 173. Krungkrai SR, Wutipraditkul N, Krungkrai J. Dihydroorotase of human malarial parasite Plasmodium falciparum differs from host enzyme. Biochem Biophys Res Commun. 2008;366(3):821-826.
  174. 174. Krungkrai SR, Prapunwattana P, Horii T, Krungkrai J. Orotate phosphoribosyltransferase and orotidine 5'-monophosphate decarboxylase exist as multienzyme complex in human malaria parasite Plasmodium falciparum. Biochem Biophys Res Commun. 2004;318(4):1012-1018.
  175. 175. Krungkrai SR, DelFraino BJ, Smiley JA, Prapunwattana P, Mitamura T, Horii T, et al. A novel enzyme complex of orotate phosphoribosyltransferase and orotidine 5'-monophosphate decarboxylase in human malaria parasite Plasmodium falciparum: physical association, kinetics, and inhibition characterization. Biochemistry. 2005;44(5):1643-1652.
  176. 176. Queen SA, Jagt DL, Reyes P. In vitro susceptibilities of Plasmodium falciparum to compounds which inhibit nucleotide metabolism. Antimicrob Agents Chemother. 1990;34(7):1393-1398.
  177. 177. Rathod PK, Khatri A, Hubbert T, Milhous WK. Selective activity of 5-fluoroorotic acid against Plasmodium falciparum in vitro. Antimicrob Agents Chemother. 1989;33(7):1090-1094.
  178. 178. Scott HV, Gero AM, O'Sullivan WJ. In vitro inhibition of Plasmodium falciparum by pyrazofurin, an inhibitor of pyrimidine biosynthesis de novo. Mol Biochem Parasitol. 1986;18(1):3-15.
  179. 179. Suttle DP, Stark GR. Coordinate overproduction of orotate phosphoribosyltransferase and orotidine-5'-phosphate decarboxylase in hamster cells resistant to pyrazofurin and 6-azauridine. J Biol Chem. 1979;254(11):4602-4607.
  180. 180. Zhang Y, Evans GB, Clinch K, Crump DR, Harris LD, Fröhlich RF, et al. Transition state analogues of Plasmodium falciparum and human orotate phosphoribosyltransferases. J Biol Chem. 2013;288(48):34746-34754.
  181. 181. Zhang Y, Luo M, Schramm VL. Transition states of Plasmodium falciparum and human orotate phosphoribosyltransferases. J Am Chem Soc. 2009;131(13):4685-4694.
  182. 182. Zhang Y, Schramm VL. Pyrophosphate interactions at the transition states of Plasmodium falciparum and human orotate phosphoribosyltransferases. J Am Chem Soc. 2010;132(25):8787-8794.
  183. 183. Kumar S, Krishnamoorthy K, Mudeppa DG, Rathod PK. Structure of Plasmodium falciparum orotate phosphoribosyltransferase with autologous inhibitory protein-protein interactions. Acta Crystallogr F Struct Biol Commun. 2015;71(Pt 5):600-608.
  184. 184. Miller BG, Wolfenden R. Catalytic proficiency: the unusual case of OMP decarboxylase. Annu Rev Biochem. 2002;71:847-885.
  185. 185. Langley DB, Shojaei M, Chan C, Lok HC, Mackay JP, Traut TW, et al. Structure and inhibition of orotidine 5'-monophosphate decarboxylase from Plasmodium falciparum. Biochemistry. 2008;47(12):3842-3854.
  186. 186. Bello AM, Poduch E, Fujihashi M, Amani M, Li Y, Crandall I, et al. A potent, covalent inhibitor of orotidine 5'-monophosphate decarboxylase with antimalarial activity. J Med Chem. 2007;50(5):915-921.
  187. 187. Bello AM, Poduch E, Liu Y, Wei L, Crandall I, Wang X, et al. Structure-activity relationships of C6-uridine derivatives targeting plasmodia orotidine monophosphate decarboxylase. J Med Chem. 2008;51(3):439-448.
  188. 188. Takashima Y, Mizohata E, Krungkrai SR, Fukunishi Y, Kinoshita T, Sakata T, et al. The in silico screening and X-ray structure analysis of the inhibitor complex of Plasmodium falciparum orotidine 5'-monophosphate decarboxylase. J Biochem. 2012;152(2):133-138.
  189. 189. Kanchanaphum P, Krungkrai J. Kinetic benefits and thermal stability of orotate phosphoribosyltransferase and orotidine 5'-monophosphate decarboxylase enzyme complex in human malaria parasite Plasmodium falciparum. Biochem Biophys Res Commun. 2009;390(2):337-341.
  190. 190. Ali MH, Imperiali B. Protein oligomerization: how and why. Bioorg Med Chem. 2005;13(17):5013-5020.
  191. 191. Levy ED, Boeri Erba E, Robinson CV, Teichmann SA. Assembly reflects evolution of protein complexes. Nature. 2008;453(7199):1262-1265.
  192. 192. Hashimoto K, Nishi H, Bryant S, Panchenko AR. Caught in self-interaction: evolutionary and functional mechanisms of protein homooligomerization. Phys Biol. 2011;8(3):035007.
  193. 193. Perica T, Chothia C, Teichmann SA. Evolution of oligomeric state through geometric coupling of protein interfaces. Proc Natl Acad Sci U S A. 2012;109(21):8127-8132.
  194. 194. Nishi H, Hashimoto K, Madej T, Panchenko AR. Evolutionary, physicochemical, and functional mechanisms of protein homooligomerization. Prog Mol Biol Transl Sci. 2013;117:3-24.
  195. 195. Caffrey DR, Somaroo S, Hughes JD, Mintseris J, Huang ES. Are protein-protein interfaces more conserved in sequence than the rest of the protein surface? Protein Sci. 2004;13(1):190-202.
  196. 196. Valdar WS, Thornton JM. Protein-protein interfaces: analysis of amino acid conservation in homodimers. Proteins. 2001;42(1):108-124.
  197. 197. Agrawal N, Dasaradhi PV, Mohmmed A, Malhotra P, Bhatnagar RK, Mukherjee SK. RNA interference: biology, mechanism, and applications. Microbiol Mol Biol Rev. 2003;67(4):657-685.
  198. 198. Hammond SM. Dicing and slicing: the core machinery of the RNA interference pathway. FEBS Lett. 2005;579(26):5822-5829.
  199. 199. Barnes RL, Shi H, Kolev NG, Tschudi C, Ullu E. Comparative genomics reveals two novel RNAi factors in Trypanosoma brucei and provides insight into the core machinery. PLoS Pathog. 2012;8(5):e1002678.
  200. 200. Mueller AK, Hammerschmidt-Kamper C, Kaiser A. RNAi in Plasmodium. Curr Pharm Des. 2014;20(2):278-283.
  201. 201. Kolev NG, Tschudi C, Ullu E. RNA interference in protozoan parasites: achievements and challenges. Eukaryot Cell. 2011;10(9):1156-1163.
  202. 202. Baum J, Papenfuss AT, Mair GR, Janse CJ, Vlachou D, Waters AP, et al. Molecular genetics and comparative genomics reveal RNAi is not functional in malaria parasites. Nucleic Acids Res. 2009;37(11):3788-3798.

Written By

Sergey Lunev, Fernando A. Batista, Soraya S. Bosch, Carsten Wrenger and Matthew R. Groves

Submitted: 06 May 2016 Reviewed: 07 September 2016 Published: 30 November 2016