Open access peer-reviewed chapter

Purine and Pyrimidine Pathways as Antimalarial Targets

Written By

Yacoba V. T. Minnow and Vern L. Schramm

Submitted: 25 June 2022 Reviewed: 11 July 2022 Published: 16 August 2022

DOI: 10.5772/intechopen.106468

From the Edited Volume

Malaria - Recent Advances and New Perspectives

Edited by Pier Paolo Piccaluga

Chapter metrics overview

135 Chapter Downloads

View Full Metrics

Abstract

Malaria continues to plague the endemic regions of sub-Saharan Africa and Southeast Asia. With the current development of artemisinin resistance and a risk of failure of the current first line therapies, there is a growing need for novel antimalarials. Purine and pyrimidine metabolism in Plasmodium is distinctly different from the human host, making these pathways valid targets for the development of novel antimalarials. Targeting key enzymes in these pathways with transition state analogs has provided high affinity inhibitors. Transition state mimicry can also provide selectivity for the parasite enzymes over the homologous enzymes of the human host. Resistance of Plasmodium parasites to current antimalarials will be compared to resistance development induced by transition state analogs inhibitors, a feature that may contribute to decreased resistance development. Tight binding and specificity of transition state analog inhibitors provide important features for novel antimalaria therapy with low toxicity and prevention of antibiotic resistance.

Keywords

  • purine salvage
  • pyrimidines
  • antimalarials
  • transition state analog inhibitors
  • resistance

1. Introduction

Malaria remains the leading cause of mortality in the endemic regions of Sub-Saharan Africa and Southeast Asia. The WHO estimates that in 2020, there were 241 million malaria cases globally with 627,000 associated deaths, an increase from years prior [1, 2]. Additionally, although progress continues to be made in some regions, such as Southeast Asia reporting a 78% reduction in malaria cases, Sub-Saharan Africa accounts for more than 90% of the global malaria incidence. Children 5 years and younger are highly susceptible to malaria with a 77% mortality rate [1, 3, 4].

Malaria is caused by the apicomplexan parasite, Plasmodium sp. The five main species able to cause disease in humans are Plasmodium falciparum, Plasmodium vivax, Plasmodium ovale, Plasmodium knowlesi and Plasmodium malariae. Plasmodium falciparum (P. falciparum) is the most prevalent and the most lethal of these [5]. Malaria infection is established when an infected female anopheles mosquito takes a blood meal from humans and introduces sporozoites into the bloodstream which invade human liver hepatocytes. The sporozoites replicate in the hepatocytes for 5–8 days and can generate thousands of merozoites which are released to infect the erythrocytes [6, 7]. Asexual replication occurs in the erythrocytes where parasites undergo several nuclear divisions and generate between 6 and 30 daughter merozoites per infected erythrocyte [6]. Symptomatic malaria is characterized by the sustained parasitization and destruction of red blood cells as well as the host immune responses. Thus, drug development efforts are directed toward the intra-erythrocytic asexual stage of the parasite.

Over the centuries-long battle with malaria, treatment has been developed from both natural product and chemical synthetic sources. Quinine, originally isolated from the bark of the cinchona tree, gave way to its synthetic counterparts, chloroquine, amodiaquine, and mefloquine, among others [8, 9]. Likewise, artemisinin was isolated from the Chinese herb, Artemisia annua (Qinghao), and gave way to the synthetic artemisinin derivatives now used in the first line malaria treatment, artemisinin-based combination therapy (ACT) [10]. However, resistance to these antimalaria treatments, have slowed global efforts toward the eradication of malaria [11, 12, 13, 14, 15]. Resistance development has led to increased efforts to identify novel antimalarials, including the recent development of the RTS,S malaria vaccine, providing a potential positive outlook for malaria control [16, 17]. New drug development efforts for malaria must identify novel targets, their mechanisms of action, and must be well tolerated with minimal side effects. Short treatment periods are essential to promote compliance and to minimize resistance development.

Transition state mimicry is a promising approach to malaria drug discovery. The technique utilizes features of transition state structure of essential target enzymes to produce tight binding and highly selective analogs [18]. Transition state analogs (TSAs) against the purine and pyrimidine pathway enzymes of P. falciparum have resulted in inhibitor molecules with the potential for new antimalarial combination therapies.

Rapid nucleic acid synthesis during intra-erythrocytic parasitic growth makes purine and pyrimidine synthesis in Plasmodium parasites an important target for novel drug development. Although both host and parasite share some enzymes in these pathways, key differences allow for Plasmodium-specific and selective molecules.

Advertisement

2. Purine metabolism

Plasmodium parasites are purine auxotrophs as they do not express the enzymes necessary to perform de novo purine synthesis [19, 20]. Substantial new DNA and RNA synthesis occurs during the asexual cycle of Plasmodium, requiring large amounts of purine and pyrimidine nucleotide precursors and the energetic contribution from ATP. Parasites therefore rely completely on the host erythrocytes for the salvage of purine nucleosides and nucleobases for the synthesis of adenylate and guanylate compounds. Hence the parasite genome contains highly expressed genes corresponding to purine transporters and purine salvage enzymes [21]. Human erythrocytes contain high (millimolar) concentrations of adenylate nucleotides, but Plasmodium parasites have no kinase to phosphorylate adenosine molecules or ribosyltransferases to salvage adenine from the host. Therefore, purine salvage in Plasmodium requires ATP catabolism, through hypoxanthine formation, by essential enzymes making up the purine salvage pathway (Figure 1) [22]. The central concept in targeting purine salvage pathways for Plasmodium involves prevention of hypoxanthine and inosine formation in both erythrocytes and parasites and/or prevention of hypoxanthine or inosine conversion to purine nucleotides in the parasites.

Figure 1.

Purine salvage pathway in P. falciparum and human erythrocytes. In P. falciparum, but not in humans, there is a link to the polyamine pathway via MTA.

Purine precursors formed in erythrocytes must gain entry to the parasite for conversion to nucleotides. Pathways proposed to facilitate the uptake of purine nucleosides and nucleobases across the parasite membrane include; (1) saturable or facilitated transport of adenosine, (2) non-saturable, channel like transport and (3) the tubovesicular membranes induced by parasitization of the host erythrocytes [2324]. A major path for purine uptake is known to be the equilibrative nucleoside transporters (ENTs) that promote nucleoside and nucleobase transport into the parasite. Four classes of ENTs have been identified in Plasmodium falciparum and Plasmodium vivax (PfENT 1-4, PvENT 1-4) [24, 25, 26, 27, 28]. Inhibitors designed against PfENT1 prevent the uptake of purines and inhibit the growth of parasites in culture [29, 30]. The essentiality of the purine salvage pathway to the survival of Plasmodium parasites, makes targeting the component enzymes attractive for novel antimalarial therapies. These include adenosine deaminase (ADA), purine nucleoside phosphorylase (PNP) and hypoxanthine-guanine-xanthine phosphoribosyltransferase (HGXPRT). Antimalarial validation of these targets has been confirmed by potent inhibitors that cause purine starvation and death of the parasites.

2.1 Adenosine deaminase (ADA, EC 3.5.4.4)

Plasmodium falciparum ADA (PfADA) catalyzes the zinc-dependent irreversible deamination of adenosine and deoxyadenosine to form inosine and deoxyinosine. PfADA can also catalyze the hydrolysis of 5′-methylthioadenosine (MTA) to 5′-methylthioinosine (MTI), a reaction that is distinctly absent from the mammalian ADA [31]. MTA is a product of the polyamine biosynthetic pathway, establishing that polyamine synthesis also plays a role in the purine salvage pathways for Plasmodium parasites. Parasites must metabolize MTA to prevent feedback inhibition of the polyamine biosynthetic pathway. Polyamine synthesis is also critical to the parasites since they cannot salvage polyamines from the host erythrocytes [see below]. Coformycin and 2′-deoxycoformycin are transition state or intermediate-like powerful inhibitors of ADA. They were originally identified as natural products and inhibit both human and parasite enzymes with picomolar affinities (Figure 2) [32]. 2′-Deoxycoformycin (Pentostatin) is an FDA-approved treatment for hairy cell leukemia, where the accumulation of adenosine or 2′-deoxyadenosine, leads to an unbalanced nucleotide pool, initiating apoptosis in B and T cells [32]. Unfortunately, there is insignificant inhibition of Plasmodium parasite growth by these inhibitors.

Figure 2.

Transition state analogs of HsADA and PfADA. The 5′-functionalized inhibitors provide improved selectivity for the parasite enzyme over the human enzyme. 5′-methylthio inhibitors are more potent thant the 5′-propylthio and 5′-phenylthio inhibitors.

The dual specificity of PfADAs serves to guide the synthesis of novel PfADA inhibitors with a high degree of selectivity for the parasite enzyme. In addition, the transition state structures of PfADA and human ADA (HsADA) are similar but distinct, with PfADA displaying an early transition state and a more intermediate-like transition state for HsADA, corresponding to their distinct catalytic turnover numbers (kcat) (Figure 3). These differences are reflected by the state of protonation at N1 of the purine ring and the N1-H bond distances. The differences permit the design and synthesis of Plasmodium-specific transition state analogs [33]. Earlier work reported the synthesis of 5′-methythiolcoformycin (MT-coformycin), a sub nanomolar (Ki* = 0.43 nM) transition state analog inhibitor of plasmodium ADA with over a 20,000-fold selectivity over the human ADA. Other functionalized 5′- and 2′-deoxycoformycin molecules also retained selectivity for the Plasmodium enzyme (Figure 2) [31, 32]. MT-coformycin and coformycin inhibited the growth of plasmodium parasites in culture when MTA is used as the sole purine source, clearly demonstrating the deaminase as an essential step in conversion of MTA toward hypoxanthine [31, 34]. This effect is not observed when adenosine or MTI is used as a purine source, demonstrating that an intact polyamine pathway is not required for purine salvage in Plasmodium species. The high degree of selectivity of the 5′-functionalized coformycins demonstrates the importance of utilizing substrate specificity and transition state analysis to design target-selective inhibitors.

Figure 3.

Reaction scheme and transition state structure of ADA.

2.2 Purine nucleoside phosphorylase (PNP, EC 2.4.2.1)

The enzyme responsible for forming purine bases from nucleosides in the purine salvage pathway is PNP, catalyzing the reversible phosphorolysis of inosine to hypoxanthine and ribose 1-phosphate [35]. Plasmodium PNP substrate specificity includes MTI, to generate hypoxanthine and 5-methylthioribose 1-phosphate [36]. MTI appears to be a parasite specific metabolite and has not been identified in mammalian pathways [37]. Hypoxanthine serves as a key precursor to nucleotide synthesis; therefore, the parasite adopts multiple pathways to generate hypoxanthine. PNP also converts guanosine and 2′-deoxyguanosine to form guanine and (2-deoxy) ribose 1-phosphate (Figure 1). PNP displays specificity for 6-oxopurines with adenine containing purine rings having no effect as substrates or inhibitors [31].

Plasmodium falciparum PNP (PfPNP) and human PNP (HsPNP) have distinct structures and catalytic sites and share only ~20% sequence similarity [38]. Structurally, PfPNP is more similar to PNP from bacterial sources and is homohexameric, displaying a trimer of dimer structure, while HsPNP like other mammalian PNP is trimeric (Figure 4). Like PfADA, PfPNP displays a broad substrate specificity that includes 6-oxopurine nucleosides and 5′-methylthio 6-oxopurine nucleosides as substrates. These differences form the basis for generation of species-specific inhibitors with discrimination for the Plasmodium enzyme.

Figure 4.

A, Homohexameric structure of PfPNP displaying a trimer of dimer conformation (1NW4). B, trimeric structure of HsPNP (1PF7).

The transition state structures of PfPNP and HsPNP, reveal similar catalytic mechanisms involving the formation of an oxocarbenium ion at the transition state (Figure 5). However, the two catalytic site interactions differ in the hydrogen bond and van der Waals interactions formed with the leaving group. The result is a more constrained active site environment at the transition state for PfPNP than for HsPNP [39]. The Immucillin-H (ImmH) transition state analog was designed to resemble the transition state structure of bovine PNP and inhibits that enzyme with a Ki* of 23 pM. Although there are distinct differences between the transition state structures of HsPNP, PfPNP and bovine PNP, ImmH is a potent inhibitor of all three enzymes, with HsPNP Ki* = 56 pM and PfPNP Ki* = 600 pM [40, 41]. DADMe-ImmH and DADMe-ImmG were developed as second generation PNP inhibitors designed to mimic the fully dissociative transition state, and thereby more closely resemble the transition state structures of HsPNP and PfPNP [42]. The introduction of a methylene bridge increases the bond distance between the 9-deazahypoxanthine (DADMe-ImmH) or the 9-deazaguanine (DADMe-ImmG) and the riboxocarbenium mimic. Additionally, the 9-deazapurine scaffold increases the pKa to permit protonation at N7 and the cation at N1′ mimics the cationic charge at this position of the transition state to more closely mimic the transition state. These features of the transition state translate to an improvement in the dissociation constant for HsPNP from 56 pM for ImmH to 16 pM for DADMe-ImmH and to 2 pM for DADMe-ImmG. Transition state analogs bind tightly to their cognate enzymes by converting the catalytic potential into thermodynamic binding. The kcat for HsPNP is approximately 50 times greater than for PfPNP, and this is reflected in the affinity of the inhibitors for the two enzyme species. Thus, the dissociation constant (Ki*, the Kd after slow onset inhibition) for PfPNP is 500 pM for DADMe-ImmH and is 670 pM for DADMe-ImmG, reflecting its lower catalytic potential [42, 43, 44].

Figure 5.

Reaction scheme and transition state structure of PNP (upper panel). Lower panel; transition state analogs of PNP with dissociation constants for HsPNP and PfPNP. Methylthio-Immucillin-H provides over 100-fold selectivity for the parasite enyzyme.

Inhibition of both HsPNP and PfPNP is required to cause a purine-less death of Plasmodium falciparum parasites in culture, as formation of hypoxanthine in erythrocytes or in the parasite will meet the needs for purine base salvage. In cell cultures, HsPNP inhibition occurs at low concentrations of PNP inhibitors with inhibition of PfPNP occurring at higher concentrations. Inhibition of both PNPs is thus required for the antiparasitic effects of the immucillins, with the IC50 values corresponding to the molar concentration of PNP present [45, 46]. The purine-less death induced by PNP inhibitors targets hypoxanthine production. Confirmation of hypoxanthine starvation as the mechanism of action, and the lack of other targets from this therapy, comes from the full restoration of parasite growth when hypoxanthine is added to culture media [45].

The genetic deficiency of HsPNP is known to cause a T cell immune deficiency by causing accumulation of 2′-deoxyguanosine (dGuo), a metabolite completely dependent on HsPNP for its recycling. Without HsPNP, activated T cells transport dGuo and convert it to dGTP which can reach toxic levels for DNA polymerase. In humans, activated T cells are therefore sensitive to the loss of HsPNP function with no effect observed in naive T cells, B cells, and other mammalian cells [47, 48]. The selective toxic effects of HsPNP inhibition in human T cells is slow to develop, taking a year or more to be symptomatic in newborns. The more rapid effect in Plasmodium suggests that inhibiting both human and parasite PNPs for antimalaria therapy will have minimal side effects. Furthermore, oral administration of DADMe-ImmG clears Plasmodium falciparum parasites from infected Aotus monkeys [49]. Although not yet tested in clinical trials for malaria, the companion compound, DADMe-ImmH, is also a powerful inhibitor of both Hs- and PfPNPs and has been in extensive phase 2 clinical trials for gout [50]. PNP inhibitors as an antimalarial therapy block hypoxanthine formation in both erythrocytes and parasites. Human blood hypoxanthine is present at approximately 1 μM, while Aotus blood was found to have approximately 40 μM hypoxanthine [49]. Therefore, PNP inhibition as an antimalarial therapy is expected to be a more effective in humans than in the Aotus test animal where several days of oral drug therapy were required to clear the parasites [49]. This hypothesis remains to be tested.

Other selective inhibitors remain an attractive option when developing novel antimalarial therapeutics. Therefore, using the dual specificity of PfPNP for MTI, the transition state analog, 5′-methylthio-immucillin-H (MT-ImmH) was synthesized [46]. MT-ImmH has a 100-fold specificity for PfPNP over HsPNP, underscoring the importance of exploring transition state chemistry and substrate specificity to develop selective inhibitors (Figure 5) [46].

2.3 Hypoxanthine-guanine-xanthine phosphoribosyltransferase (HGXPRT, EC 2.4.2.8, EC 2.4.2.22)

Hypoxanthine is the essential purine precursor for Plasmodium and HGXPRT is an essential enzyme in the Plasmodium purine salvage pathway, as the only enzyme to incorporate hypoxanthine into the parasite nucleotide pool. HGXPRT catalyzes the phosphoribosylation of hypoxanthine, guanine, and xanthine using 5-phosphoribosyl-α-D-1-pyrophosphate (PRPP) to generate inosine 5′-monophosphate (IMP), guanosine 5′-monophosphate (GMP) and xanthine 5′-monophosphate (XMP) respectively [51]. Pyrophosphate (PPi) is generated as the second product of the reaction (Figure 6). IMP serves as the precursor for the synthesis of both adenine and guanine-based nucleotides. The single distinction in substrate specificity between the Plasmodium HGXPRT (PfHGPRT) and the human HGPRT (HsHGPRT) is the inability of the human homolog to use xanthine as a substrate [52]. Xanthine in humans is converted to uric acid by xanthine oxidase and is the terminal end of purine metabolism, followed by its excretion into the urine [53, 54].

Figure 6.

Reaction scheme for HG(X)PRT. PfHGXPRT has a unique substrate specificity for xanthine. HsHGPRT does not.

PfHGXPRT has been identified as the most highly expressed purine salvage enzyme in P. falciparum and has been reported to have the highest activity compared to other enzymes of purine salvage identified by Reyes and coworkers [55]. Parasites may have evolved to use hypoxanthine as a key nucleotide precursor because it is the end-product of purine catabolism in human erythrocytes, with subsequent oxidation to uric acid occurring in other tissues. The high amounts of intracellular phosphates in erythrocytes favors the activity of PNP to produce hypoxanthine [55, 56]. Since its identification, PfHGXPRT has been the most targeted enzyme for purine salvage in P. falciparum. Detailed kinetic and structural analyses have led to the design of many inhibitors of PfHGXPRT activity. Challenges in targeting PfHGXPRT include its similarities to HsHGPRT and the need for the anionic phosphate group, making cell access a problem [57, 58, 59, 60].

PfHGXPRT and HsHGPRT share 76% sequence similarity and 44% sequence identity [61]. Active site residues in contact with substrates and inhibitors are conserved (Figure 7). Purine nucleoside analogs and transition state analogs have been investigated and many prove to be good inhibitors of both PfHGXPRT and HsHGPRT. Acyclic nucleoside phosphonates (ANPs) however provide the greatest selectivity for the Plasmodium enzyme compared to phosphate-based compounds (Figure 8) [62, 63]. Crystal structures of HsHGPRT in complex with ANP and free PfHGXPRT suggest that the increased in selectivity when the phosphate groups are replaced with phosphonates is attributed to tighter hydrogen bonding around the phosphonate group and flexibility of the 6-oxo binding pocket [60].

Figure 7.

Two-dimensional ligand interaction map showing A, Immucillin-G phosphate bound to HsHGPRT (1BZY) and B, Immucillin-H phosphate bound PfHGXPRT (1CJB). Active side residues that make contact with inhibitors are similar in both enzymes.

Figure 8.

First generation and second generation transtion state analogs for HG(X)PRT. Dissociation constants for HsHGPRT and PfHGXPRT are shown for each inhibitor. The second generation analogs provide greater selectivity for the PfHGXPRT. Lysophospholipid prodrug with biological activity against P. falciparum parasites in culture shown in lower panel.

TSAs are the tightest binding HG(X)PRT inhibitors reported to date. Although the first-generation transition state analogs, immucillin-G phosphate (ImmGP) and immucillin-H phosphate (ImmHP) were designed based on the proposed similarity of the transition state structure to other phosphoribosyltransferases, they bind potently but without discrimination for parasite or human HG(X)PRT (Figure 8) [41] and are subject to host enzymatic degradation by phosphomonoesterases. Structural analysis of TSAs bound to both HsHGPRT and PfHGXPRT as well as downfield proton chemical shift differences observed by NMR of bound inhibitors suggested that selectivity for PfHGXPRT may be achieved by substituting TSAs with electron-withdrawing or electron-donating groups to alter hydrogen-bonding distances [41]. Recently, the transition state structure of PfHGXPRT has been solved and provides detailed information on bond distances in the transition state [61]. With this knowledge, serinol-based mimics of the riboxocarbenium with a methylene bridge linking the oxocarbenium mimic to the 9-deazapurine ring were designed to mimic the bond distances in the transition state structure and synthesized. These compounds have proven to be the tightest and most selective inhibitors of PfHGXPRT, with Ki* values as low as 650 pM and selectivity indices for PfHGXPRT greater than 500-fold relative to HsHGXPRT [64, 65].

HGXPRT TSAs are phosphate or phosphonate-based compounds with a net negative charge at physiological pH, making them impermeable to cell membranes. Prodrug approaches are therefore required for their biological activity. Prodrugs of PfHGXPRT inhibitors involve ANPs with intramolecular esters in the form of lysophospholipid prodrugs, designed to be activated by intracellular phospholipase C, or phosphoramidate prodrugs with a multistep activation mechanism [66]. Prodrug approaches have been validated in several FDA-approved nucleoside antiviral drugs [67]. The prodrug approaches improve the biological activities of TSAs of PfHGXPRT to give anti-parasite IC50 values in the micromolar range (2–7 μM) against P. falciparum parasites cultured in human blood (Figure 8) [65]. With a Ki* of 650 pM, disparity between the Ki* for the enzyme and the micromolar IC50 values of lysophospholipid prodrugs suggests that more efficient prodrug approaches are needed to fully capture the tight-binding capabilities of TSAs intracellularly. PfHGXPRT is a challenging target because of the multiple cellular and membrane barriers between an oral drug and the intracellular parasites. However, similar challenges have been overcome for antivirals and they provide a lesson for prodrug approaches. In summary, PfHGXPRT is a challenging but valid target for the development of novel antimalarials and TSAs currently provide the most promising approach.

2.4 Downstream enzymes of purine salvage

Given how essential purine salvage is to the survival of Plasmodium parasites, it is remarkable that the pathway evolved to rely on only three major enzymes, ADA, PNP and HGXPRT, based on the genome interpretation and protein expression levels in the parasites [55]. As concentration of free guanine in human blood is very low, it is not feasible for Plasmodium spp. to synthesize guanosine monophosphate (GMP) using HGXPRT [51]. Two enzymes downstream of HGXPRT, inosine 5′-monophosphate dehydrogenase (IMPDH) and GMP synthetase (GMPS) work sequentially to synthesize GMP (Figure 1) [68]. These two enzymes control GMP synthesis in most organisms [69]. IMPDH catalyzes the NAD+ dependent reaction that coverts IMP to xanthine monophosphate (XMP) and NADH [70]. GMPS is composed of two domains, an ATP pyrophosphatase domain and a glutamine amidotransferase (GAT) domain [68, 71]. By this mechanism, GMPS catalyzes the ATP-dependent irreversible amination of XMP on carbon 2 to form GMP. Inhibitors of both IMPDH and GMPS have been shown to have antimalaria properties and have been the subject of recent drug development [21, 68].

Mycophenolic acid is an IMPDH inhibitor and inhibits P. falciparum parasites in culture with an IC50 of about 5 μM [21, 72, 73]. Bredinin is an inhibitor that has been shown to inhibit both IMPDH and GMPS [68, 74] and suggest that both IMPDH and GMPS are targets to explore for novel antimalarials. Although the IMP to GMP pathway involves phosphorylated intermediates, both mycophenolic acid and bredinin are non-phosphorylated natural products (or derivatives) that provide orally available access to tissues. Both mycophenolic acid and bredinin are used in immune suppression, and they have not been useful as antimalarials.

2.5 Purine salvage and polyamine synthesis

Polyamines are synthesized by the transfer of propyl amino groups from decarboxylated S-adenosylmethionine to putrescine to form spermidine (one transfer) and spermine (two transfers) [75]. MTA is the product of each transfer in the polyamine biosynthetic pathway. MTA is also a substrate for PfADA, whereby it is deaminated to methylthioinosine. These steps provide a path for this abundant product of the polyamine pathway to be channeled into the into the purine salvage pathway [76]. Polyamines are present at high amounts in the intraerythrocytic stages of Plasmodium spp., which can perform both de novo synthesis and salvage of polyamines [76, 77]. Given that parasites can only perform salvage of purines, the valuable purines used in the polyamine pathway are economically recycled. MTA is converted by the action of PfADA to produce MTI. MTI is a substrate for PfPNP to form hypoxanthine and methylthioribose 1-phosphate. MTI is not found in human metabolism and recent studies have suggested that MTI production by Plasmodium falciparum in infected humans leads to activation of human Toll-like receptor 8 (TLR8), a signal for the stimulation of host innate immunity [37]. This connection provides an interesting link to the potential use of PfPNP inhibitors as a potential therapeutic in human malaria. MTI is removed from P. falciparum metabolism exclusively by PfPNP as HsPNP does not use MTI as substrate. Therefore, inhibition of purine salvage via PfPNP will increase MTI, increase signaling from TLR8 and have an immune stimulatory effect for the clearance of infected erythrocytes. Crosstalk between the polyamine pathway and purine salvage in Plasmodium parasites, together with the potential immunostimulatory effect of polyamine metabolites, underscores the importance of the polyamine pathway to purine salvage in Plasmodium parasites [37]. Inhibition of either the polyamine or the purine salvage pathways in Plasmodium parasites will be detrimental to parasites by; (1) purine starvation and death and (2) immune stimulation to promote clearance by host immune cells.

Advertisement

3. Pyrimidine metabolism

Opposite to the purine requirements of Plasmodium falciparum, the parasites are incapable of pyrimidine salvage from host erythrocytes; and pyrimidine synthesis requires the de novo pathway. This requirement contrasts with host cells with pathways for de novo synthesis and salvage of pyrimidine nucleotides [78]. Pyrimidine synthesis occurs by the action of six sequential enzymes to produce uridine monophosphate (UMP) as the primary product and a precursor to the other pyrimidine nucleotides (Figure 9) [79]. The enzymatic steps following UMP synthesis in P. falciparum are not well characterized, however, pyrimidine biosynthesis is linked to folate metabolism via the action of thymidylate synthetase (TS) [80]. Differences in the pathway between the human host and Plasmodium are important for selective targeting of Plasmodium parasites. In humans, the first three enzymes of the pyrimidine biosynthetic pathway, carbamoylphosphate synthetase II (CPS II), aspartate transcarbamoylase (ATC) and dihydroorotase (DHO) are a single protein, multiple domains, multifunctional enzyme complex, the CAD complex. In Plasmodium, CPS II, ATC and DHO are encoded as distinct genes, produce independent proteins and are organized on different locations on chromosomes, Ch. 13 for CPS II and ATC, and Ch. 14 for DHO [78, 81].

Figure 9.

De novo pyrimidine biosynthetic pathway. The pyrimidine biosynthetic pathway is conserved in Plasmodium and humans.

CPS II and ATC in P. falciparum are poorly characterized despite the role of CPS II as a key regulator of the pathway [20, 82]. The most thoroughly characterized enzymes of Plasmodium pyrimidine biosynthesis are dihydroorotate dehydrogenase (DHODH) and orotate phosphoribosyltransferase (OPRT) [83]. The transition state structures of human (HsOPRT) and Plasmodium falciparum OPRT (PfOPRT) have previously been solved and transition state analog inhibitors of both HsOPRT and PfOPRT have been designed (detailed below).

3.1 DHODH (EC 1.3.5.2)

DHODH is the fourth enzyme in the pyrimidine biosynthetic pathway and is expressed as two isozymes. The isozyme expressed from chromosome 7 codes for a mitochondria-associated DHODH and chromosome 9 expresses a cytosolic DHODH, where the mitochondrial form is essential for Plasmodium growth [84]. DHODH catalyzes the formation of orotate from dihydroorotate, a rate-limiting step for pyrimidine biosynthesis. Plasmodium spp., like the human host, express a type II DHODH, localized in the inner membrane of the mitochondrion [84, 85]. The electron transport chain provides quinone for the redox activity of DHODH. Recent studies have highlighted PfDHODH as a potential antimalaria target, showing that inhibiting PfDHODH leads to parasite death in culture and animal models [85, 86, 87].

Selective inhibition of PfDHODH has been achieved with DSM265, a triazolopyrimidine with a 4000-fold selectivity over mammalian DHODH. DSM265 has resulted in phase II clinical trials for the treatment of malaria, a first for any inhibitor of the parasite pyrimidine biosynthetic pathway. DSM265 inhibits both blood and liver stages of the malaria parasite with a biological efficacy similar to chloroquine in murine malaria models [87, 88, 89]. The pharmacokinetic analysis of DSM265 supported single dose therapy for use as a once weekly prophylactic [87]. Single dose administrations are considered important for patient compliance. Since the discovery of DSM265, DSM421, an improvement on DSM265 in terms of compound solubility, plasma stability and equal targeting of P. falciparum and P. vivax has also been reported in preclinical development [90]. Other studies have followed suit, using pharmacophore screening and structure-guided virtual studies to identify plasmodium specific DHODH inhibitors with submicromolar IC50 values [91, 92].

Pyrimidine biosynthesis is coupled to the mitochondrial electron transport chain (ETC) by DHODH and its requirements of quinones for catalytic activity (Figure 9). This coupling results in indirect inhibition of DHODH by ETC inhibitors such as atovaquone, an approved antimalarial often used in combination with proguanil [93]. Although DHODH and the linked mitochondrial ETC were promising targets with extensive discovery and development programs providing powerful inhibitors, the rapid development of resistance in clinical trials has hindered their development toward approved drugs [94, 95].

3.2 OPRT (EC 2.4.2.10)

Reduction of dihydroorotate by DHODH produces orotic acid and the following step, catalyzed by PfOPRT, catalyzes the production of the nucleotide orotidine monophosphate (OMP) from orotate and PRPP. OMP is subsequently converted to UMP by OMP decarboxylase (OMPDC). Plasmodium and human OPRTs differ in the early steps of the pyrimidine biosynthetic pathway. Thus, human OPRT is fused to OMPDC to form the single protein, bifunctional enzyme called UMP synthase [96]. In Plasmodium, OPRT and OPMDC have been purified as monofunctional enzymes however, there are recent reports suggesting that OPRT and OPMDC may exist as a heterotetrameric enzyme in Plasmodium with the potential for hydrolysis into the single enzymes during purification [83, 97, 98]. Detailed kinetic characterizations of both human and P. falciparum OPRTs (HsOPRT and PfOPRT respectively) have been reported and active site differences between the homologs have been characterized [83, 98].

The essential function of OPRT in pyrimidine biosynthesis and therefore for parasite proliferation has been demonstrated by the selective killing of cultured P. falciparum parasites by pyrazofurin, an inhibitor of OPRT with anti-parasite IC50 values of 6–20 μM [21, 99]. Additionally, 5-fluororotate and 5-aminouracil have inhibitory activity against PfOPRT with IC50 values of 42 nM and 8 μM respectively [21]. 5-Fluororotate selectively targets P. falciparum parasites in culture with an IC50 of 6 nM [100]. The mechanism of action is proposed to arise from an indirect effect of the toxic 5-fluorodeoxyuridylate metabolite on TS [101]. Plasmodium express TS as a single protein, multi-domain bifunctional enzyme with dihydrofolate reductase (DHFR) whereas TS exists as a monofunctional enzyme in mammals [21, 99, 100].

Transition state analysis of PfOPRT and HsOPRT reported an SN1 transition state with a partial dissociative orotate and riboxocarbenium ion character for both enzymes [102]. Despite differences in the enzyme organization, PfOPRT and HsOPRT possess similar transition state structures, where the ribocation is fully developed, the dianionic orotate is fully dissociated and there is weak participation of the pyrophosphate nucleophile (Figure 10) [102]. The transition state structures provided information for the design of several TSAs with nanomolar affinities for both HsOPRT and PfOPRT (Figure 10) [103]. Inhibition of HsOPRT might have potential utility for the treatment of autoimmune diseases and some malignant neoplastic diseases. However, no anti-parasitic activity was observed for any of the TSAs in cultured P. falciparum. The lack of activity was attributed to permeability barriers and target access, but no permeability studies have been reported [103]. PfOPRT remains a viable antimalarial target requiring further studies to elucidate and develop species-specific antimalarials.

Figure 10.

Upper panel: OPRT reaction scheme and transition state structure. The transition state structure is similar for PfOPRT and HsOPRT. It features a fully developed ribocation and a fully dissociated dianionic orotate with weak participation of the pyrophosphate. Lower panel: Transition state analogs synthesized and tested on PfOPRT and HsOPRT.

3.3 OMPDC (EC 4.1.1.23)

The decarboxylation of OMP to form UMP is catalyzed by OMPDC, the sixth enzyme in the de novo pyrimidine biosynthetic pathway [78]. OMPDC has achieved catalytic notoriety as it has been considered the most proficient pure protein catalyst, giving a catalytic reaction rate enhancement of approximately 1017 [104, 105]. Unlike other decarboxylases, OMPDC requires no metal ions or cofactors for its catalytic activity [105]. Recent studies indicate that Plasmodium falciparum OMPDC may form a tight heterotetrameric complex with OPRT (PfOMPDC2-PfOPRT2) with properties distinct from the bifunctional human OPRT [106, 107]. Although both are expressed and encoded by two separate genes, unique amino acid insertions in both proteins not present in the homologous proteins of other organisms allow the formation of the heterotetrameric complex [98, 106]. Additional support comes from the kinetic characterization of fused P. falciparum OPMDC and OPRT, expressed in Escherichia coli, where the catalytic efficiency of the fused enzymes was enhanced several orders of magnitude relative to either enzyme acting as monofunctional proteins, leading the authors to call the fusion a ‘super’ enzyme [97].

Nucleoside 5′-monophosphate analogs, 6-azauridine 5′-monophosphate (AzaUMP), allopurinol-3-riboside 5′-monophosphate (allopurinol MP), pyrazofurin 5′-monophosphate (pyrazofuin MP), and xanthosine 5′-monophosphate (XMP) are potent inhibitors of both human and parasite OMPDCs (Figure 11) [108]. These display a strong preference for PfOMPDC, with selectivity ranging from 11 to 150-fold preference for the parasite enzyme [108]. The differential binding of XMP has been investigated by comparing X-ray crystal structures of both enzymes. Few differences in active site residues were observed and specificity may be as a result in hydrogen bonding differences between ligand and active site residues (Figure 12). Another reported difference is the βα5-loop present at the dimerization interface, which displays different conformations and amino acid substitutions in PfOMPDC compared to HsOMPDC and is postulated to cause different active site rearrangements around the pyrimidine binding region [108]. The phosphate binding loop is also larger in HsOMPDC than in PfOMPDC [108]. Repositioning of the loop is a proposed requirement to bind XMP as observed for XMP-bound Pyrococcus horikoshii OMPDC [108]. Interestingly, the phosphate loop is the same size and shape in P. horikoshii OMPDC and PfOMPDC, therefore, these enzymes may bind XMP similarly. XMP is a 150-fold more potent inhibitor of PfOMPDC than of HsOMPDC [108].

Figure 11.

Nucleoside 5′-analogs of OPMDC showing inhibitory constants for both PfOMPDC and HsOMPDC. Inhibitors are highly selective, 11–150-fold for PfOMPDC over HsOMPDC. Values are as reported in [108].

Figure 12.

2-dimensional ligand interaction map of UMP bound A, PfOPMDC (6DSR) and B, HsOMPDC (2V30). Similar active site residues make contact with UMP.

No crystal structures of inhibitor bound PfOMPDC are available to understand the basis for inhibitor selectivity. However, the high preference of inhibitors for parasite OMPDC compared to the human enzyme indicates the possibility of developing potent and selective antimalarials. Inhibition of OMPDC has focused on nucleotide analogs, where the 5′-phosphate is a significant contributor to binding. A substantial barrier to drug development for nucleotide inhibitors is the anionic charge, as phosphoryl anions are membrane impermeable and phosphomonoesters are susceptible to phosphomonoesterases. Prodrug approaches, now being used in antiviral therapy for the delivery of phosphate nucleosides may be useful in development of OMPDC inhibitors.

3.4 DHFR (EC 1.5.1.3) and TS (EC 2.1.1.45)

De novo purine synthesis in the host and pyrimidine synthesis in both host and parasites are dependent on folate metabolites for the donation of carbon units. The pyrimidine biosynthetic pathway is linked to folate metabolism by TS, which requires 5,10-methylene tetrahydrofolate as a methyl donor for its catalytic activity [109]. DHFR and TS are both targets for human cancers [110]. In Plasmodium falciparum, DHFR and TS are expressed as a single protein, two-domain bifunctional DHFR-TS enzyme system, catalyzing the synthesis of deoxythymidylate monophosphate (dTMP) from dUMP and the NADPH-dependent reduction of dihydrofolate (DHF) to tetrahydrofolate (THF) [80]. The bifunctional coupling of adjacent steps in the pathway is proposed to cause “substrate channeling”, an efficient utilization of folate production for the synthesis of purines, pyrimidines, and amino acids [111, 112]. Since malaria parasites are unable to salvage pyrimidines, if metabolites or cofactors required by the folate pathway are cut off, thymidine biosynthesis is blocked, and parasites die from pyrimidine starvation [20, 80]. Thus, DHFR has been a frequently-drugged target for antimalaria therapy.

Pyrimethamine (PYR) and proguanil (PG) are approved antimalaria therapeutics which selectively target PfDHFR by several hundred-fold compared to the human counterparts [20, 113]. Mutations in the DHFR gene conferring resistance to both PYR and PG have led to diminished clinical effectiveness as single therapeutics for malaria. However, the combination of pyrimethamine with sulfadoxine (SP), which blocks dihydropteroate synthase, a folate precursor, act synergistically in the treatment of severe malaria [114]. Despite the prevalence of molecular mutations causing resistance to SP, WHO still recommends SP for intermittent preventative treatment for pregnant women and, often, in combination with amodiaquine, an inhibitor of Plasmodium heme polymerase, for seasonal malaria prevention [115]. PG is also used in combination with atovaquone, an inhibitor of the mitochondrion electron transfer chain, which also directly impacts pyrimidine biosynthesis because of its coupling to DHODH [116, 117].

Recent searches for additional inhibitors of DHFR-TS identified the antimicrobial triclosan to inhibit the proliferation of intraerythrocytic P. falciparum and is a 775 nM inhibitor of P. vivax DHFR. Triclosan displays a 20-fold selectivity for Plasmodium enzymes relative to human DHFR [118]. To preempt the development of resistance, Tarnchompoo and coworkers have developed hybrid inhibitors of DHFR with both flexible and rigid side chains that target both wild-type and multiple resistance P. falciparum [119]. Hybrid inhibitors include sub-nanomolar inhibitors of PfDHFR with 10-fold selectivity over HsDHFR [119]. Flexible cycloguanil analogs have been characterized that are low nanomolar inhibitors of multiple mutant PfDHFR [120]. Fragment-based screening approaches to inhibitor design have identified non-pyrimidine scaffolds that inhibit PfDHFR-TS with IC50 in the range of 28–695 μM and are highly selective over HsDHFR [121]. This approach offers a new avenue to tackle antifolate resistance and develop new antifolate antimalarials, but remains in development [121].

DHFR-TS is a validated target for antimalaria combination therapy by blocking synthesis of dTMP. The functional characteristics of DHFR-TS including its crystal structure and the structures of drug-resistant DHFR-TS allows for the identification and development of next-generation selective inhibitors of either catalytic domain that will address resistance to current antifolate drugs for malaria. New approaches are being explored, including peptide-based antagonists of PfDHFR-TS in the quest for novel and potent antifolate/anti-pyrimidine drugs [122].

Advertisement

4. Transition state analogs and drug development

Transition state theory postulates that chemically stable mimics of the enzymatic transition state will bind tightly to the target [123, 124]. All chemical, including enzymatic reactions, proceed via a transition state, a transient high energy species that lies along the reaction coordinate between reactants and products [125, 126]. Transition states are the balance point of a catalytic reaction where bonds are partially broken or formed and the probability for product formation and return to reactants is equal. Enzymes have evolved to highly favor the transition state geometry, hence their very large rate enhancement factors, typically of 1010 to 1015, a very large decrease in the activation energy of the reaction [18]. Chemically stable mimics of the transition state bind to and stabilize this favored enzymatic transition state geometry, capture part of the transition state energy and bind very tightly [18, 127, 128, 129]. Therefore, transition state analysis provides a powerful tool to develop molecules with high inhibitory potential for their target enzymes. Kinetic isotope effects (KIE) remain the best approach to study enzymatic transition states and together with quantum computational chemistry, allows the construction of electrostatic potential maps that provide information on the transition state structure [18]. This information enables the design of transition state analogs that bind with high fidelity to their cognate enzymes, millions of times tighter than substrates. Transition state analogs are some of the tightest binding enzymatic inhibitors and have the propensity to bind their target enzymes with dissociation constants in the nanomolar to femtomolar range (10−9 to 10−15 M) [123, 125, 126]. A feature of transition state analogs is the ability to convert the enzyme potential for catalysis into binding energy, therefore the more catalytically efficient the enzyme, the tighter the potential for transition state analog binding [127].

The immucillins are chemically stable transition state analogs that mimic the ribocation transition state of N-ribosyltransferases, enzymes that include PfPNP and PfHGXPRT, both important in purine salvage in P. falciparum [50]. Immucillin-H (also known as: BCX1777, Forodesine and Mundesine) (Figure 5) is an inhibitor of PNPs and has been approved in Japan for the treatment of resistant or relapsed peripheral T cell lymphoma (PTCL). Immucillin-A, as Galidesivir (also known as BCX4430) continues in phase I - II clinical trials for antiviral therapy. It is converted to the triphosphate form where it blocks RNA chain elongation in RNA viruses, including Yellow Fever and SARS-CoV2 [130]. DADMe-Immucillin-H (also known as: BCX4208, Ulodesine) has completed phase II clinical trials for gout, by virtue of its powerful inhibition of human PNP, an essential step in formation of uric acid in humans [50]. No immucillins or related transition state analogs for purine and pyrimidine pathways have yet entered clinical trials for antimalaria therapy though several are in preclinical testing [128].

The immucillins differ from traditional antibiotic discovery. Antibacterials are often discovered by cell wall screening or genomic targeting to inhibit cell wall synthesis [131]. Immucillins and related transition state analogs are designed to mimic the geometry of the transition state of their specific enzymes [129]. The resulting transition state analogs can be powerful tight-binding inhibitors with exquisite specificity for their targeted enzymes. Transition state inhibitor design is adding new candidates for drug development.

Immucillins in preclinical development as antimalaria drugs include the PNP-targeting DADMe-ImmG, a powerful inhibitor that is a picomolar inhibitor for both human and PfPNPs. Administration at nanomolar concentrations to parasites cultured in human erythrocytes caused purine-less death of P. falciparum. Testing in Aotus primates infected with P. falciparum at an oral dose of 50 mg/kg and dosed for 7 days gave robust parasite clearance and no parasitemia was detected by day 6 of treatment, along with complete inhibition of host and parasite PNP [49]. Recrudescent parasites appeared several days after treatment stopped [49]. Aotus monkeys, however, have approximately 40 μM circulating hypoxanthine, compared to approximately 1 μM in humans [49]. Therefore, clearing P. falciparum from Aotus is a more stringent test for the antimalaria efficacy of DADMe-ImmG than anticipated in human infections. DADMe-ImmH, the compound tested extensively in phase I and phase II clinical trials for gout, has a good safety profile in humans. DADMe-ImmH has an inhibitory potential for both HsPNP and PfPNPs similar to DADMe-ImmG and is a strong candidate to enter human clinical trials [50]. The efficacy of DADMe-ImmH has been established in clinical trials. A single oral dose of 0.5 mg/kg completely inhibits erythrocyte PNP with inhibition lasting for 120 days, the lifetime of the erythrocyte, making it a candidate for single-dose therapy of P. falciparum [132]. As inhibition of both human PNP and PfPNP is required for antimalarial properties, a single dose therapy has potential as a novel antimalarial, and as described below, has a reduced potential for the induction of resistance mutations.

4.1 Antimalaria resistance to transition state analogs

P. falciparum has developed resistance to most approved antimalarials, contributing to the difficulty of disease-eradication efforts worldwide. Rapid development of resistance has led to the clinical failure of several potent antimalaria drugs early in clinical trials and has resulted in World Health Organization recommendations for combination drug therapy for malaria treatment. Chloroquine which was once the gold standard treatment for malaria was widely used until the 1950s, when widespread resistance prompted its removal from the list of approved malaria therapeutics [133, 134]. Resistance soon followed for drugs such as atovaquone, pyrimethamine, proguanil, cycloguanil, sulfadoxine and sulfadoxine-pyrimethamine (SP) [133, 134]. Recently, resistance to artemisinin, one partner of the current first line treatment has been reported in the endemic regions of Sub-Saharan Africa and the Greater Mekong region of Southeast Asia [12, 14, 135, 136].

In the field, drug resistance is characterized by delayed-clearance of parasites from the blood of infected individuals [11]. Resistant parasites often have point mutations and gene amplification which can result in decreased drug uptake, increased efflux of drugs, target overexpression, target modification that reduces drug binding, and inactivation of a drug by modification (Figure 13) [134]. Knowledge of these mechanisms to resistance have prompted the search for novel and unique antimalaria compounds followed by investigation of the rates and mechanisms of resistance by subjecting cultured P. falciparum parasites to constant pressure of the novel compounds.

Figure 13.

Antimalarial resistance mechanism frequently employed by Plasmodium parasites to overcome drug pressure.

One such study involving DSM265, the PfDHODH inhibitor. Using a drug pressure of 3 × EC50 revealed a minimum inoculum for resistance (MIR) of 2 × 106 compared to 2 × 107 for atovaquone [87]. Selection of resistant parasites to DSM265 required 8 x EC50 of DSM265 compared to 33 × EC50 required to suppress atovaquone resistant parasites. The resistance mechanism in Dd2 clones (resistant to chloroquine, pyrimethamine and mefloquine) revealed both gene amplification of the DHODH gene and the G181C mutation that confers a 13-fold shift in IC50 of the drug compared to wild-type enzyme and 26-fold reduction of the EC50 in parasite growth assay. Kinetic characterization of the recombinant mutant enzyme revealed a 2-fold increase in the kcat, and Km was unchanged [87]. Resistance to DSM265 in Dd2 parasites developed rapidly with a profile similar to resistance development for atovaquone. However, resistance was not as easily generated at higher concentrations of DSM265 in other P. falciparum strains, questioning if resistance would develop in the field [90]. DSM265 in phase II clinical trials for malaria led to the appearance of resistance during the clinical trial phase [94].

In contrast to DSM265, resistance to transition state analogs is not easily generated. Loss of transition state features in the protein also generates severe catalytic penalties predicted to induce a fitness cost to the parasites. For example, treatment of cultured P. falciparum with DADMe-ImmG for 1 year in culture resulted in a 6-fold resistance (6-fold increase in IC50) [44]. The resistance was caused by a 6-fold amplification of wild-type PfPNP gene and protein levels. After 3 years of drug pressure (over 2136 clonal selections) in cultured cells, increased resistance was marked by a 12-fold amplification in the target PfPNP gene [44]. The 12-fold gene amplification was accompanied by two point mutations in PNP occurring in separate clones (M183L and V181D) to give rise to a 500-fold increase in the IC50. Interestingly, point mutations were only present in 50% of the transcripts and the remainders were wild type [44]. Recombinant expression of mutant PNP revealed that the kinetic properties were incompatible with the purine salvage function of PNP [44]. Particularly, the M183L mutation resulted in a 17,000-fold decrease in the catalytic efficiency (30-fold decrease in kcat) of the enzyme and a 39,000-fold decrease in affinity for DADMe-ImmG [44]. Therefore, the mutation reduced the catalytic efficiency and DADMe-ImmG efficacy by approximately the same amount, highlighting a key feature of transition state analogs. Mutations that prevent binding of the analog are expected to decrease catalytic activity to the same degree.

Resistance to DADMe-ImmG in P. falciparum arises by a unique mechanism. Hybrid PNP expression of 50% mutant M183L subunits and 50% wild-type subunits is proposed to generate hybrid hexameric PNPs demonstrating a 6-fold decrease in catalytic efficiency, as three of the six subunits are native. This hybrid hexameric construct of three native and three mutant subunits displays negative cooperativity in binding to DADMe-ImmG, always leaving a fraction of the native subunits free to catalyze the formation of hypoxanthine [137]. Therefore, in addition to 12-fold target overexpression and point mutations, P. falciparum also employs hybrid multimeric PNP to achieve robust resistance [137]. This example emphasizes the severe catalytic costs and threat to biological function that resistance to transition state analogs produces in the parasites. This slow development of resistance to DADMe-ImmG suggests that resistance may be slow to develop in the field.

Conditions for generating laboratory DADMe-ImmG-resistant P. falciparum clones are stringent, requiring up to 3 years of drug pressure on the same culture before robust resistance appeared. In the field, effective anti-parasitics clear parasites in a matter of days, and it is unlikely that parasites would encounter conditions of continuous drug pressure. Notwithstanding, these in vitro resistance assays help identify mechanisms and molecular markers of resistance and resistance phenotypes that will aid in recognizing resistance development in the field. Like most antimalarials, the proposed use of DADMe-ImmG or related transition state analogs would be in drug combinations to provide an extra layer of protection against development of resistance.

Advertisement

5. Summary and conclusion

During the intraerythrocytic stages, Plasmodium parasites proliferate rapidly and require extensive nucleic acid synthesis. The building blocks are either salvaged from the human host in the form of purines or synthesized de novo (pyrimidines). There are distinct differences between the purine and pyrimidine synthesis pathways of humans and the malaria parasites. Therefore, targeting these pathways can aid in the development of novel chemotherapeutic agents to combat emerging resistance to the current first line antimalarials. Transition state analogs against various purine and pyrimidine pathway enzymes are emerging as promising candidates for antimalarial therapy. The slow development of resistance and the unusual resistance mechanism employed by the parasites to drug pressure against transition state analogs may lead to slow resistance development in the field. Finally, the recent advances made with the RTS,S vaccine in endemic African countries is promising and point in the right direction for population control of malaria. However, until an effective vaccine is approved for widespread use, chemotherapeutic agents managing the disease and vector control measures are still needed.

Advertisement

Acknowledgments

Preparation of this manuscript was supported by NIH research grants (AI127807 and GM041916). We thank our collaborators in Dr. Peter C. Tyler’s group at Ferrier Research Institute, Victoria University of Wellington, New Zealand, for the synthesis of the Immucillins and transition state analogs.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. WHO. World Malaria Report 2021. Geneva: World Health Organization; 2021
  2. 2. WHO. World Malaria Report 2018. Geneva: World Health Organization; 2018
  3. 3. WHO. World Malaria Report 2020. Geneva: World Health Organization; 2020
  4. 4. WHO. World Malaria Report. Geneva: World Health Organization; 2019
  5. 5. Sato S. Plasmodium-a brief introduction to the parasites causing human malaria and their basic biology. Journal of Physiological Anthropology. 2021;40(1):1. DOI: 10.1186/s40101-020-00251-9
  6. 6. White NJ, Pukrittayakamee S, Hien TT, Faiz MA, Mokuolu OA, Dondorp AM. Malaria. Lancet. 2014;383(9918):723-735. DOI: 10.1016/S0140-6736(13)60024-0
  7. 7. Haldar K, Bhattacharjee S, Safeukui I. Drug resistance in Plasmodium. Nature Reviews. Microbiology. 2018;16(3):156-170. DOI: 10.1038/nrmicro.2017.161
  8. 8. Achan J et al. Quinine, an old anti-malarial drug in a modern world: Role in the treatment of malaria. Malaria Journal. 2011;10:144. DOI: 10.1186/1475-2875-10-144
  9. 9. Renslo AR. Antimalarial drug discovery: From quinine to the dream of eradication. ACS Medicinal Chemistry Letters. 2013;4(12):1126-1128. DOI: 10.1021/ml4004414
  10. 10. Miller LH, Su X. Artemisinin: Discovery from the Chinese herbal garden. Cell. 2011;146(6):855-858. DOI: 10.1016/j.cell.2011.08.024
  11. 11. Tilley L, Straimer J, Gnadig NF, Ralph SA, Fidock DA. Artemisinin action and resistance in plasmodium falciparum. Trends in Parasitology. 2016;32(9):682-696. DOI: 10.1016/j.pt.2016.05.010
  12. 12. Dondorp AM et al. Artemisinin resistance: Current status and scenarios for containment. Nature Reviews. Microbiology. 2010;8(4):272-280. DOI: 10.1038/nrmicro2331
  13. 13. Thu AM, Phyo AP, Landier J, Parker DM, Nosten FH. Combating multidrug-resistant Plasmodium falciparum malaria. The FEBS Journal. 2017;284(16):2569-2578. DOI: 10.1111/febs.14127
  14. 14. Balikagala B et al. Evidence of Artemisinin-Resistant Malaria in Africa. The New England Journal of Medicine. 2021;385(13):1163-1171. DOI: 10.1056/NEJMoa2101746
  15. 15. Imwong M et al. The spread of artemisinin-resistant plasmodium falciparum in the greater Mekong subregion: A molecular epidemiology observational study. The Lancet Infectious Diseases. 2017;17(5):491-497. DOI: 10.1016/S1473-3099(17)30048-8
  16. 16. Laurens MB. RTS,S/AS01 vaccine (Mosquirix): An overview. Human Vaccines & Immunotherapeutics. 2020;16(3):480-489. DOI: 10.1080/21645515.2019.1669415
  17. 17. Zavala F. RTS,S: The first malaria vaccine. Journal of Clinical Investigation. 2022;132(1). DOI: 10.1172/JCI156588
  18. 18. Schramm VL. Enzymatic transition states and drug design. Chemical Reviews. 2018;118(22):11194-11258. DOI: 10.1021/acs.chemrev.8b00369
  19. 19. Downie MJ, Kirk K, Mamoun CB. Purine salvage pathways in the intraerythrocytic malaria parasite Plasmodium falciparum. Eukaryotic Cell. 2008;7(8):1231-1237. DOI: 10.1128/EC.00159-08
  20. 20. Hyde JE. Targeting purine and pyrimidine metabolism in human apicomplexan parasites. Current Drug Targets. 2007;8(1):31-47. DOI: 10.2174/138945007779315524
  21. 21. Queen SA, Jagt DL, Reyes P. In vitro susceptibilities of Plasmodium falciparum to compounds which inhibit nucleotide metabolism. Antimicrobial Agents and Chemotherapy. 1990;34(7):1393-1398. DOI: 10.1128/AAC.34.7.1393
  22. 22. Berman PA, Human L, Freese JA. Xanthine oxidase inhibits growth of Plasmodium falciparum in human erythrocytes in vitro. The Journal of Clinical Investigation. 1991;88(6):1848-1855. DOI: 10.1172/JCI115506
  23. 23. Piro F, Focaia R, Dou Z, Masci S, Smith D, Di Cristina M. An uninvited seat at the dinner table: How apicomplexan parasites scavenge nutrients from the host. Microorganisms. 2021;9(12):2592. DOI: 10.3390/ microorganisms9122592
  24. 24. Carter NS et al. Isolation and functional characterization of the PfNT1 nucleoside transporter gene from Plasmodium falciparum. The Journal of Biological Chemistry. 2000;275(14):10683-10691. DOI: 10.1074/jbc.275.14.10683
  25. 25. Quashie NB, Ranford-Cartwright LC, de Koning HP. Uptake of purines in Plasmodium falciparum-infected human erythrocytes is mostly mediated by the human equilibrative nucleoside transporter and the human facilitative nucleobase transporter. Malaria Journal. 2010;9:36. DOI: 10.1186/1475-2875-9-36
  26. 26. Frame IJ, Merino EF, Schramm VL, Cassera MB, Akabas MH. Malaria parasite type 4 equilibrative nucleoside transporters (ENT4) are purine transporters with distinct substrate specificity. Biochemical Journal. 2012;446(2):179-190. DOI: 10.1042/BJ20112220
  27. 27. Kirk K, Lehane AM. Membrane transport in the malaria parasite and its host erythrocyte. Biochemical Journal. 2014;457(1):1-18. DOI: 10.1042/BJ20131007
  28. 28. Downie MJ, Saliba KJ, Howitt SM, Broer S, Kirk K. Transport of nucleosides across the Plasmodium falciparum parasite plasma membrane has characteristics of PfENT1. Molecular Microbiology. 2006;60(3):738-748. DOI: 10.1111/j.1365-2958.2006.05125.x
  29. 29. Frame IJ et al. Yeast-based high-throughput screen identifies Plasmodium falciparum equilibrative nucleoside transporter 1 inhibitors that kill malaria parasites. ACS Chemical Biology. 2015;10(3):775-783. DOI: 10.1021/cb500981y
  30. 30. Deniskin R, Frame IJ, Sosa Y, Akabas MH. Targeting the plasmodium vivax equilibrative nucleoside transporter 1 (PvENT1) for antimalarial drug development. International Journal for Parasitology: Drugs and Drug Resistance. 2016;6(1):1-11. DOI: 10.1016/j.ijpddr.2015.11.003
  31. 31. Ting LM et al. Targeting a novel Plasmodium falciparum purine recycling pathway with specific immucillins. The Journal of Biological Chemistry. 2005;280(10):9547-9554. DOI: 10.1074/jbc.M412693200
  32. 32. Tyler PC, Taylor EA, Frohlich RF, Schramm VL. Synthesis of 5′-methylthio coformycins: Specific inhibitors for malarial adenosine deaminase. Journal of the American Chemical Society. 2007;129(21):6872-6879. DOI: 10.1021/ja0708363
  33. 33. Luo M, Singh V, Taylor EA, Schramm VL. Transition-state variation in human, bovine, and Plasmodium falciparum adenosine deaminases. Journal of the American Chemical Society. 2007;129(25):8008-8017. DOI: 10.1021/ja072122y
  34. 34. Ho MC et al. Structural and metabolic specificity of methylthiocoformycin for malarial adenosine deaminases. Biochemistry. 2009;48(40):9618-9626. DOI: 10.1021/bi9012484
  35. 35. Bzowska A, Kulikowska E, Shugar D. Purine nucleoside phosphorylases: Properties, functions, and clinical aspects. Pharmacology & Therapeutics. 2000;88(3):349-425. DOI: 10.1016/s0163-7258(00)00097-8
  36. 36. Donaldson TM et al. Structural determinants of the 5′-methylthioinosine specificity of plasmodium purine nucleoside phosphorylase. PLoS ONE. 2014;9(1):e84384. DOI: 10.1371/journal.pone.0084384
  37. 37. Köllisch G et al. TLR8 is activated by 5′-methylthioinosine, a Plasmodium falciparum-derived intermediate of the purine salvage pathway. Cell Reports. 2022;39(2):110691. DOI: 10.1016/j. celrep.2022.110691
  38. 38. Kicska GA, Tyler PC, Evans GB, Furneaux RH, Kim K, Schramm VL. Transition state analogue inhibitors of purine nucleoside phosphorylase from Plasmodium falciparum. The Journal of Biological Chemistry. 2002;277(5):3219-3225. DOI: 10.1074/jbc.M105905200
  39. 39. Lewandowicz A, Schramm VL. Transition state analysis for human and Plasmodium falciparum purine nucleoside phosphorylases. Biochemistry. 2004;43(6):1458-1468. DOI: 10.1021/bi0359123
  40. 40. Ho MC et al. Four generations of transition-state analogues for human purine nucleoside phosphorylase. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(11):4805-4812. DOI: 10.1073/pnas.0913439107
  41. 41. Li CM et al. Transition-state analogs as inhibitors of human and malarial hypoxanthine-guanine phosphoribosyltransferases. Nature Structural Biology. 1999;6(6):582-587. DOI: 10.1038/9367
  42. 42. Evans GB, Furneaux RH, Lewandowicz A, Schramm VL, Tyler PC. Synthesis of second-generation transition state analogues of human purine nucleoside phosphorylase. Journal of Medicinal Chemistry. 2003;46(24):5271-5276. DOI: 10.1021/jm030305z
  43. 43. Evans GB, Furneaux RH, Tyler PC, Schramm VL. Synthesis of a transition state analogue inhibitor of purine nucleoside phosphorylase via the Mannich reaction. Organic Letters. 2003;5(20):3639-3640. DOI: 10.1021/ol035293q
  44. 44. Ducati RG et al. Genetic resistance to purine nucleoside phosphorylase inhibition in Plasmodium falciparum. Proceedings of the National Academy of Sciences of the United States of America. 2018;115(9):2114-2119. DOI: 10.1073/pnas.1525670115
  45. 45. Kicska GA, Tyler PC, Evans GB, Furneaux RH, Schramm VL, Kim K. Purine-less death in Plasmodium falciparum induced by immucillin-H, a transition state analogue of purine nucleoside phosphorylase. The Journal of Biological Chemistry. 2002;277(5):3226-3231. DOI: 10.1074/jbc.M105906200
  46. 46. Madrid DC, Ting LM, Waller KL, Schramm VL, Kim K. Plasmodium falciparum purine nucleoside phosphorylase is critical for viability of malaria parasites. The Journal of Biological Chemistry. 2008;283(51):35899-35907. DOI: 10.1074/jbc.M807218200
  47. 47. Bantia S, Kilpatrick JM. Purine nucleoside phosphorylase inhibitors in T-cell malignancies. Current Opinion in Drug Discovery & Development. 2004;7(2):243-247. [Online]. Available from: https://www.ncbi.nlm.nih.gov/pubmed/15603259
  48. 48. Kicska GA et al. Immucillin H, a powerful transition-state analog inhibitor of purine nucleoside phosphorylase, selectively inhibits human T lymphocytes. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:4593-4598, 4598. DOI: 10.1073/pnas.071050798
  49. 49. Cassera MB et al. Plasmodium falciparum parasites are killed by a transition state analogue of purine nucleoside phosphorylase in a primate animal model. PLoS ONE. 2011;6(11):e26916. DOI: 10.1371/journal.pone.0026916
  50. 50. Evans GB, Tyler PC, Schramm VL. Immucillins in infectious diseases. ACS Infectious Diseases. 2018;4(2):107-117. DOI: 10.1021/acsinfecdis.7b00172
  51. 51. Cheviet T, Lefebvre-Tournier I, Wein S, Peyrottes S. Plasmodium purine metabolism and its inhibition by nucleoside and nucleotide analogues. Journal of Medicinal Chemistry. 2019;62(18):8365-8391. DOI: 10.1021/acs.jmedchem.9b00182
  52. 52. Keough DT, Ng AL, Winzor DJ, Emmerson BT, de Jersey J. Purification and characterization of Plasmodium falciparum hypoxanthine-guanine-xanthine phosphoribosyltransferase and comparison with the human enzyme. Molecular and Biochemical Parasitology. 1999;98(1):29-41. DOI: 10.1016/s0166-6851(98)00139-x
  53. 53. Moffatt BA, Ashihara H. Purine and pyrimidine nucleotide synthesis and metabolism. Arabidopsis Book. 2002;1:e0018. DOI: 10.1199/tab. 0018
  54. 54. Leung HB, Schramm VL. The role of adenosine monophosphate nucleosidase in the regulation of adenine nucleotide levels in Azotobacter vinelandii during aerobic-anaerobic transitions. Archives of Biochemistry and Biophysics. 1978;190(1):46-56. DOI: 10.1016/0003-9861(78)90252-7
  55. 55. 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. Molecular and Biochemical Parasitology. 1982;5(5):275-290. DOI: 10.1016/0166-6851(82)90035-4
  56. 56. Furuhashi M. New insights into purine metabolism in metabolic diseases: Role of xanthine oxidoreductase activity. American Journal of Physiology. Endocrinology and Metabolism. 2020;319(5):E827-E834. DOI: 10.1152/ajpendo.00378.2020
  57. 57. Roy S, Nagappa LK, Prahladarao VS, Balaram H. Kinetic mechanism of Plasmodium falciparum hypoxanthine-guanine-xanthine phosphoribosyltransferase. Molecular and Biochemical Parasitology. 2015;204(2):111-120. DOI: 10.1016/j.molbiopara.2016.02.006
  58. 58. Roy S, Karmakar T, Prahlada Rao VS, Nagappa LK, Balasubramanian S, Balaram H. Slow ligand-induced conformational switch increases the catalytic rate in plasmodium falciparum hypoxanthine guanine xanthine phosphoribosyltransferase. Molecular Biosystems. 2015;11(5):1410-1424. DOI: 10.1039/c5mb00136f
  59. 59. Makobongo MO et al. The purine salvage enzyme hypoxanthine guanine xanthine phosphoribosyl transferase is a major target antigen for cell-mediated immunity to malaria. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(5):2628-2633. DOI: 10.1073/pnas.0337629100
  60. 60. Keough DT et al. Inhibition of hypoxanthine-guanine phosphoribosyltransferase by acyclic nucleoside phosphonates: A new class of antimalarial therapeutics. Journal of Medicinal Chemistry. 2009;52(14):4391-4399. DOI: 10.1021/jm900267n
  61. 61. Ducati RG, Firestone RS, Schramm VL. Kinetic isotope effects and transition state structure for hypoxanthine-guanine-xanthine phosphoribosyltransferase from Plasmodium falciparum. Biochemistry. 2017;56(48):6368-6376. DOI: 10.1021/acs.biochem.7b01027
  62. 62. Keough DT et al. Acyclic nucleoside phosphonates containing a second phosphonate group are potent inhibitors of 6-oxopurine phosphoribosyltransferases and have antimalarial activity. Journal of Medicinal Chemistry. 2013;56(6):2513-2526. DOI: 10.1021/jm301893b
  63. 63. Keough DT et al. Aza-acyclic nucleoside phosphonates containing a second phosphonate group as inhibitors of the human, Plasmodium falciparum and vivax 6-oxopurine phosphoribosyltransferases and their prodrugs as antimalarial agents. Journal of Medicinal Chemistry. 2015;58(2):827-846. DOI: 10.1021/jm501416t
  64. 64. Clinch K et al. Acyclic phosph(on)ate inhibitors of Plasmodium falciparum hypoxanthine-guanine-xanthine phosphoribosyltransferase. Bioorganic & Medicinal Chemistry. 2013;21(17):5629-5646. DOI: 10.1016/j.bmc.2013.02.016
  65. 65. Hazleton KZ et al. Acyclic immucillin phosphonates: Second-generation inhibitors of Plasmodium falciparum hypoxanthine-guanine-xanthine phosphoribosyltransferase. Chemistry & Biology. 2012;19(6):721-730. DOI: 10.1016/j.chembiol.2012.04.012
  66. 66. Hockova D et al. Antimalarial activity of prodrugs of N-branched acyclic nucleoside phosphonate inhibitors of 6-oxopurine phosphoribosyltransferases. Bioorganic & Medicinal Chemistry. 2015;23(17):5502-5510. DOI: 10.1016/j.bmc.2015.07.038
  67. 67. Bhatia HK, Singh H, Grewal N, Natt NK. Sofosbuvir: A novel treatment option for chronic hepatitis C infection. Journal of Pharmacology and Pharmacotherapeutics. 2014;5(4):278-284. DOI: 10.4103/0976-500X.142464
  68. 68. Bhat JY, Shastri BG, Balaram H. Kinetic and biochemical characterization of Plasmodium falciparum GMP synthetase. Biochemical Journal. 2008;409(1):263-273. DOI: 10.1042/BJ20070996
  69. 69. Hedstrom L. IMP dehydrogenase: Structure, mechanism, and inhibition. Chemical Reviews. 2009;109(7):2903-2928. DOI: 10.1021/cr900021w
  70. 70. Fotie J. Inosine 5'-monophosphate dehydrogenase (IMPDH) as a potential target for the development of a new generation of Antiprotozoan agents. Mini Reviews in Medicinal Chemistry. 2018;18(8):656-671. DOI: 10.2174/1389557516666160620065558
  71. 71. Ballut L et al. Active site coupling in Plasmodium falciparum GMP synthetase is triggered by domain rotation. Nature Communications. 2015;6:8930. DOI: 10.1038/ncomms9930
  72. 72. Veletzky L et al. In vitro activity of immunosuppressive drugs against Plasmodium falciparum. Malaria Journal. 2014;13:476. DOI: 10.1186/1475-2875-13-476
  73. 73. Mogire RM et al. Target-similarity search using Plasmodium falciparum proteome identifies approved drugs with anti-malarial activity and their possible targets. PLoS ONE. 2017;12(10):e0186364. DOI: 10.1371/journal.pone.0186364
  74. 74. Webster HK, Whaun JM. Antimalarial properties of bredinin. Prediction based on identification of differences in human host-parasite purine metabolism. The Journal of Clinical Investigation. 1982;70(2):461-469. DOI: 10.1172/jci110636
  75. 75. Park MH, Igarashi K. Polyamines and their metabolites as diagnostic markers of human diseases. Biomolecules & Therapeutics. 2013;21(1):1-9. DOI: 10.4062/biomolther.2012.097
  76. 76. El Bissati K et al. Novel synthetic polyamines have potent antimalarial activities in vitro and in vivo by decreasing intracellular spermidine and spermine concentrations. Frontiers in Cellular and Infection Microbiology. 2019;9:9. DOI: 10.3389/fcimb.2019.00009
  77. 77. Ramya TN, Surolia N, Surolia A. Polyamine synthesis and salvage pathways in the malaria parasite Plasmodium falciparum. Biochemical and Biophysical Research Communications. 2006;348(2):579-584. DOI: 10.1016/j.bbrc.2006.07.127
  78. 78. Krungkrai J. Malaria parasite pyrimidine nucleotide metabolism: A promising drug target. Archives of Parasitology. 2017;1(1):101
  79. 79. Krungkrai SR, Krungkrai J. Insights into the pyrimidine biosynthetic pathway of human malaria parasite Plasmodium falciparum as chemotherapeutic target. Asian Pacific Journal of Tropical Medicine. 2016;9(6):525-534. DOI: 10.1016/j.apjtm.2016.04.012
  80. 80. Hyde JE. Exploring the folate pathway in Plasmodium falciparum. Acta Tropica. 2005;94(3):191-206. DOI: 10.1016/j.actatropica.2005.04.002
  81. 81. Moreno-Morcillo M, Grande-Garcia A, Ruiz-Ramos A, Del Cano-Ochoa F, Boskovic J, Ramon-Maiques S. Structural insight into the core of CAD, the multifunctional protein leading de novo pyrimidine biosynthesis. Structure. 2017;25(6):912-923 e5. DOI: 10.1016/j.str.2017.04.012
  82. 82. Asai T, O'Sullivan WJ, Kobayashi M, Gero AM, Yokogawa M, Tatibana M. Enzymes of the de novo pyrimidine biosynthetic pathway in Toxoplasma gondii. Molecular and Biochemical Parasitology. 1983;7(2):89-100. DOI: 10.1016/0166-6851(83)90037-3
  83. 83. Krungkrai SR et al. Human malaria parasite orotate phosphoribosyltransferase: Functional expression, characterization of kinetic reaction mechanism and inhibition profile. Molecular and Biochemical Parasitology. 2004;134(2):245-255. DOI: 10.1016/j.molbiopara.2003.12.006
  84. 84. Sykes DB. The emergence of dihydroorotate dehydrogenase (DHODH) as a therapeutic target in acute myeloid leukemia. Expert Opinion on Therapeutic Targets. 2018;22(11):893-898. DOI: 10.1080/14728222.2018.1536748
  85. 85. Singh A, Maqbool M, Mobashir M, Hoda N. Dihydroorotate dehydrogenase: A drug target for the development of antimalarials. European Journal of Medicinal Chemistry. 2017;125:640-651. DOI: 10.1016/j.ejmech.2016.09.085
  86. 86. White J et al. Identification and mechanistic understanding of dihydroorotate dehydrogenase point mutations in Plasmodium falciparum that confer in vitro resistance to the clinical candidate DSM265. ACS Infectious Diseases. 2019;5(1):90-101. DOI: 10.1021/acsinfecdis.8b00211
  87. 87. Phillips MA et al. A long-duration dihydroorotate dehydrogenase inhibitor (DSM265) for prevention and treatment of malaria. Science Translational Medicine. 2015;7(296):296ra111. DOI: 10.1126/scitranslmed.aaa6645
  88. 88. Gujjar R et al. Lead optimization of aryl and aralkyl amine-based triazolopyrimidine inhibitors of Plasmodium falciparum dihydroorotate dehydrogenase with antimalarial activity in mice. Journal of Medicinal Chemistry. 2011;54(11):3935-3949. DOI: 10.1021/jm200265b
  89. 89. Baldwin J et al. High-throughput screening for potent and selective inhibitors of Plasmodium falciparum dihydroorotate dehydrogenase. The Journal of Biological Chemistry. 2005;280(23):21847-21853. DOI: 10.1074/jbc.M501100200
  90. 90. Phillips MA et al. A triazolopyrimidine-based dihydroorotate dehydrogenase inhibitor with improved drug-like properties for treatment and prevention of malaria. ACS Infectious Diseases. 2016;2(12):945-957. DOI: 10.1021/acsinfecdis.6b00144
  91. 91. Dickerman BK et al. Identification of inhibitors that dually target the new permeability pathway and dihydroorotate dehydrogenase in the blood stage of Plasmodium falciparum. Scientific Reports. 2016;6:37502. DOI: 10.1038/srep37502
  92. 92. 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. Journal of Chemical Information and Modeling. 2016;56(3):548-562. DOI: 10.1021/acs.jcim.5b00680
  93. 93. Guler JL, White J 3rd, Phillips MA, Rathod PK. Atovaquone tolerance in Plasmodium falciparum parasites selected for high-level resistance to a dihydroorotate dehydrogenase inhibitor. Antimicrobial Agents and Chemotherapy. 2015;59(1):686-689. DOI: 10.1128/AAC.02347-14
  94. 94. Llanos-Cuentas A et al. Antimalarial activity of single-dose DSM265, a novel plasmodium dihydroorotate dehydrogenase inhibitor, in patients with uncomplicated Plasmodium falciparum or Plasmodium vivax malaria infection: A proof-of-concept, open-label, phase 2a study. The Lancet Infectious Diseases. 2018;18(8):874-883. DOI: 10.1016/S1473-3099(18)30309-8
  95. 95. Luth MR, Winzeler EA. SnapShot: Antimalarial drugs. Cell. 2020;183(2):554-554 e1. DOI: 10.1016/j.cell.2020.09.006
  96. 96. Suchi M et al. Molecular cloning of the human UMP synthase gene and characterization of point mutations in two hereditary orotic aciduria families. American Journal of Human Genetics. 1997;60(3):525-539. [Online]. Available from: https://www.ncbi.nlm.nih.gov/pubmed/9042911; https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1712531/pdf/ajhg00003-0063.pdf
  97. 97. Paojinda P, Imprasittichai W, Krungkrai SR, Palacpac NMQ , Horii T, Krungkrai J. Bifunctional activity of fused Plasmodium falciparum orotate phosphoribosyltransferase and orotidine 5′-monophosphate decarboxylase. Parasitology International. 2018;67(1):79-84. DOI: 10.1016/j.parint.2017.04.003
  98. 98. 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. Biochemical and Biophysical Research Communications. 2004;318(4):1012-1018. DOI: 10.1016/j.bbrc.2004.04.124
  99. 99. Scott HV, Gero AM, O'Sullivan WJ. In vitro inhibition of Plasmodium falciparum by pyrazofurin, an inhibitor of pyrimidine biosynthesis de novo. Molecular and Biochemical Parasitology. 1986;18(1):3-15. DOI: 10.1016/0166-6851(86)90045-9
  100. 100. Rathod PK, Khatri A, Hubbert T, Milhous WK. Selective activity of 5-fluoroorotic acid against Plasmodium falciparum in vitro. Antimicrobial Agents and Chemotherapy. 1989;33(7):1090-1094. DOI: 10.1128/AAC.33.7.1090
  101. 101. Rathod PK, Leffers NP, Young RD. Molecular targets of 5-fluoroorotate in the human malaria parasite, Plasmodium falciparum. Antimicrobial Agents and Chemotherapy. 1992;36(4):704-711. DOI: 10.1128/AAC.36.4.704
  102. 102. Zhang Y, Luo M, Schramm VL. Transition states of Plasmodium falciparum and human orotate phosphoribosyltransferases. Journal of the American Chemical Society. 2009;131(13):4685-4694. DOI: 10.1021/ja808346y
  103. 103. Zhang Y et al. Transition state analogues of Plasmodium falciparum and human orotate phosphoribosyltransferases. The Journal of Biological Chemistry. 2013;288(48):34746-34754. DOI: 10.1074/jbc.M113.521955
  104. 104. Miller BG, Wolfenden R. Catalytic proficiency: The unusual case of OMP decarboxylase. Annual Review of Biochemistry. 2002;71:847-885. DOI: 10.1146/annurev.biochem.71.110601.135446
  105. 105. Callahan BP, Miller BG. OMP decarboxylase—An enigma persists. Bioorganic Chemistry. 2007;35(6):465-469. DOI: 10.1016/j.bioorg.2007.07.004
  106. 106. Imprasittichail W, Roytrakul S, Krungkrai SR, Krungkrail J. A unique insertion of low complexity amino acid sequence underlies protein-protein interaction in human malaria parasite orotate phosphoribosyltransferase and orotidine 5′-monophosphate decarboxylase. Asian Pacific Journal of Tropical Medicine. 2014;7(3):184-192. DOI: 10.1016/S1995-7645(14)60018-3
  107. 107. Krungkrai SR 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. DOI: 10.1021/bi048439h
  108. 108. Langley DB et al. Structure and inhibition of orotidine 5′-monophosphate decarboxylase from Plasmodium falciparum. Biochemistry. 2008;47(12):3842-3854. DOI: 10.1021/bi702390k
  109. 109. Shuvalov O, Petukhov A, Daks A, Fedorova O, Vasileva E, Barlev NA. One-carbon metabolism and nucleotide biosynthesis as attractive targets for anticancer therapy. Oncotarget. 2017;8(14):23955-23977. DOI: 10.18632/oncotarget.15053
  110. 110. Antosiewicz A et al. Human dihydrofolate reductase and thymidylate synthase form a complex in vitro and co-localize in normal and cancer cells. Journal of Biomolecular Structure & Dynamics. 2017;35(7):1474-1490. DOI: 10.1080/07391102.2016.1186560
  111. 111. Meek TD, Garvey EP, Santi DV. Purification and characterization of the bifunctional thymidylate synthetase-dihydrofolate reductase from methotrexate-resistant Leishmania tropica. Biochemistry. 1985;24(3):678-686. DOI: 10.1021/bi00324a021
  112. 112. Chaianantakul N, Sirawaraporn R, Sirawaraporn W. Insights into the role of the junctional region of Plasmodium falciparum dihydrofolate reductase-thymidylate synthase. Malaria Journal. 2013;12(91). DOI: 10.1186/1475-2875-12-91
  113. 113. Reynolds MG, Roos DS. A biochemical and genetic model for parasite resistance to antifolates. Toxoplasma gondii provides insights into pyrimethamine and cycloguanil resistance in Plasmodium falciparum. The Journal of Biological Chemistry. 1998;273(6):3461-3469. DOI: 10.1074/jbc.273.6.3461
  114. 114. Chulay JD, Watkins WM, Sixsmith DG. Synergistic antimalarial activity of pyrimethamine and sulfadoxine against Plasmodium falciparum in vitro. The American Journal of Tropical Medicine and Hygiene. 1984;33(3):325-330. DOI: 10.4269/ajtmh.1984.33.325
  115. 115. Deloron P, Bertin G, Briand V, Massougbodji A, Cot M. Sulfadoxine/pyrimethamine intermittent preventive treatment for malaria during pregnancy. Emerging Infectious Diseases. 2010;16(11):1666-1670. DOI: 10.3201/eid1611.101064
  116. 116. Sibley CH et al. Pyrimethamine-sulfadoxine resistance in Plasmodium falciparum: What next? Trends in Parasitology. 2001;17(12):582-588. DOI: 10.1016/s1471-4922(01)02085-2
  117. 117. Srivastava IK, Vaidya AB. A mechanism for the synergistic antimalarial action of atovaquone and proguanil. Antimicrobial Agents and Chemotherapy. 1999;43(6):1334-1339. DOI: 10.1128/AAC.43.6.1334
  118. 118. Bilsland E et al. Plasmodium dihydrofolate reductase is a second enzyme target for the antimalarial action of triclosan. Scientific Reports. 2018;8(1):1038. DOI: 10.1038/s41598-018-19549-x
  119. 119. Tarnchompoo B et al. Hybrid inhibitors of malarial dihydrofolate reductase with dual binding modes that can forestall resistance. ACS Medicinal Chemistry Letters. 2018;9(12):1235-1240. DOI: 10.1021/acsmedchemlett.8b00389
  120. 120. Kamchonwongpaisan S et al. Flexible diaminodihydrotriazine inhibitors of Plasmodium falciparum dihydrofolate reductase: Binding strengths, modes of binding and their antimalarial activities. European Journal of Medicinal Chemistry. 2020;195:112263. DOI: 10.1016/j.ejmech.2020.112263
  121. 121. Hoarau M, Vanichtanankul J, Srimongkolpithak N, Vitsupakorn D, Yuthavong Y, Kamchonwongpaisan S. Discovery of new non-pyrimidine scaffolds as Plasmodium falciparum DHFR inhibitors by fragment-based screening. Journal of Enzyme Inhibition and Medicinal Chemistry. 2021;36(1):198-206. DOI: 10.1080/14756366.2020.1854244
  122. 122. Chaianantakul N, Sungkapong T, Supatip J, Kingsang P, Kamlaithong S, Suwanakitti N. Antimalarial effect of cell penetrating peptides derived from the junctional region of Plasmodium falciparum dihydrofolate reductase-thymidylate synthase. Peptides. 2020;131:170372. DOI: 10.1016/j.peptides.2020.170372
  123. 123. Schramm VL, Horenstein BA, Kline PC. Transition state analysis and inhibitor design for enzymatic reactions. Journal of Biological Chemistry. 1994;269(28):18259-18262. [Online]. Available from: https://www.ncbi.nlm.nih.gov/pubmed/8034566; http://www.jbc.org/content/269/28/18259.full.pdf
  124. 124. Schramm VL. Enzymatic transition states and transition state analog design. Annual Review of Biochemistry. 1998;67:693-720. DOI: 10.1146/annurev.biochem.67.1.693
  125. 125. Schramm VL. Enzymatic transition states and transition state analogues. Current Opinion in Structural Biology. 2005;15(6):604-613. DOI: 10.1016/j.sbi.2005.10.017
  126. 126. Schramm VL. Enzymatic transition states: Thermodynamics, dynamics and analogue design. Archives of Biochemistry and Biophysics. 2005;433(1):13-26. DOI: 10.1016/j.abb.2004.08.035
  127. 127. Schramm VL. Transition states and transition state analogue interactions with enzymes. Accounts of Chemical Research. 2015;48(4):1032-1039. DOI: 10.1021/acs.accounts.5b00002
  128. 128. Schramm VL. Transition states, analogues, and drug development. ACS Chemical Biology. 2013;8(1):71-81. DOI: 10.1021/cb300631k
  129. 129. Schramm VL. Enzymatic transition state theory and transition state analogue design. The Journal of Biological Chemistry. 2007;282(39):28297-28300. DOI: 10.1074/jbc.R700018200
  130. 130. Julander JG et al. An update on the progress of galidesivir (BCX4430), a broad-spectrum antiviral. Antiviral Research. 2021;195:105180. DOI: 10.1016/j.antiviral.2021.105180
  131. 131. Thomson JM, Lamont IL. Nucleoside analogues as antibacterial agents. Frontiers in Microbiology. 2019;10:952. DOI: 10.3389/fmicb.2019.00952
  132. 132. Gebre ST, Cameron SA, Li L, Babu YS, Schramm VL. Intracellular rebinding of transition-state analogues provides extended in vivo inhibition lifetimes on human purine nucleoside phosphorylase. The Journal of Biological Chemistry. 2017;292(38):15907-15915. DOI: 10.1074/jbc.M117.801779
  133. 133. Tse EG, Korsik M, Todd MH. The past, present and future of anti-malarial medicines. Malaria Journal. 2019;18(1):93. DOI: 10.1186/s12936-019-2724-z
  134. 134. Klein EY. Antimalarial drug resistance: A review of the biology and strategies to delay emergence and spread. International Journal of Antimicrobial Agents. 2013;41(4):311-317. DOI: 10.1016/j.ijantimicag.2012.12.007
  135. 135. Dondorp AM et al. Artemisinin resistance in Plasmodium falciparum malaria. The New England Journal of Medicine. 2009;361(5):455-467. DOI: 10.1056/NEJMoa0808859
  136. 136. Taylor SM, Juliano JJ, Meshnick SR. Artemisinin resistance in Plasmodium falciparum malaria. The New England Journal of Medicine. 2009;361(18):1807. DOI: 10.1056/NEJMc091737. author reply 1808
  137. 137. Minnow YVT, Harijan RK, Schramm VL. A resistant mutant of plasmodium falciparum purine nucleoside phosphorylase uses wild-type neighbors to maintain parasite survival. The Journal of Biological Chemistry. 2021;296:100342. DOI: 10.1016/j.jbc.2021.100342

Written By

Yacoba V. T. Minnow and Vern L. Schramm

Submitted: 25 June 2022 Reviewed: 11 July 2022 Published: 16 August 2022