Design of inhibitors of ODCase - Future Science

2 downloads 0 Views 2MB Size Report
ODCase is a highly proficient enzyme responsible for the decarboxylation of orotidine monophosphate to generate uridine monophosphate. ODCase has ...
Review For reprint orders, please contact [email protected]

Design of inhibitors of ODCase ODCase is a highly proficient enzyme responsible for the decarboxylation of orotidine monophosphate to generate uridine monophosphate. ODCase has attracted early attention due to its interesting mechanism of catalysis. In order to exploit therapeutic advantages due to the inhibition of ODCase, one must have selective inhibitors of this enzyme from the pathogen, or a dysregulated molecular mechanism involving ODCase. ODCase inhibitors have potential applications as anticancer agents, antiviral agents, antimalarial agents and potentially act against other parasitic diseases. A variety of C6-substituted uridine monophosphate derivatives have shown excellent inhibition of ODCase. 6-iodouridine is a potent inhibitor of the malaria parasite, and its monophosphate form covalently inhibits ODCase. A variety of inhibitors of ODCase with potential applications as therapeutic agents are discussed in this review. ODCase (EC 4.1.1.23), an enzyme discovered almost 50 years ago, was initially viewed only in the context of nucleotide biosynthesis, where it catalyzes the final reaction in the de novo biosynthesis of uridine monophosphate (UMP) (Figure 1) [1]. In the mid-1990s, however, it was discovered that ODCase is able to catalyze the decarboxylation of orotidine monophosphate (OMP) at extremely high rates compared with the conversion in its absence, popularizing the enzyme and leading to an increased interest in this enzyme [2,3]. ODCase is one of the most proficient enzymes known today [4]. The enzyme enhances the catalytic decarboxylation of OMP (1) to UMP (2) over 17 orders of magnitude compared with the conversion seen in water at neutral pH and ambient temperature, and is able to do so with no assistance from cofactors or metal ions [1,5]. In bacteria and parasites, ODCase is a monofunctional enzyme [6,7], while in humans it is part of the bifunctional enzyme, UMP synthase. ODCase’s sophisticatedly evolved mechanism of catalysis for decarboxylation has the enzyme a highly investigated subject [8,9]. The catalytic decarboxylation of OMP to UMP is a critical process within a cell. The triphosphate form of UMP is a precursor for the synthesis of RNA, as well as a precursor for the synthesis of other pyrimidine nucleotides. Pyrimidine nucleotides are also important building blocks for the synthesis of DNA. Thus, in most species, including humans, these pyrimidine nucleotides are obtained by two pathways, the de novo pathway and the salvage pathway. ODCase has a central role in the synthesis of pyrimidines as part of the de novo pathway and is therefore present in most species including

bacteria, parasites and higher level species, but not in viruses. Viruses depend on their host cells to supply them with nucleotides [10]. However, certain parasitic organisms, such as Plasmodia, are dependent solely on the de novo synthesis to obtain pyrimidine nucleotides because their cells are not equipped to utilize the salvage pathway [11]. When there is a high demand for pyrimidine nucleotides, such as during cell replication and abnormal cell growth, the de novo synthesis is upregulated [12,13]. Using ODCase inhibitors to limit or inhibit pyrimidine production in the cell could potentially elicit a variety of desirable biological activities including anti­v iral, anti­ plasmodial and anticancer activities [14]. A handful of investigations in the past focused on developing ODCase inhibitors against cancer because of the important cellular roles of activated intermediates such as UDP-glucose and CDP-choline in biosynthetic pathways. Preclinical studies suggested that the therapeutic effects of 5-fluorouracil can be enhanced by pretreatment with agents such as phosphonacetyl-l-aspartic acid (PALA), an inhibitor of aspartate transcarbamylase. The objective of treatment with PALA was to increase the activation of 5-fluorouracil by inhibiting the normal pathway of de novo pyrimidine biosynthesis. However, this combination therapy led to mild to moderate myelosuppression, mucositis, diarrhea, nausea and vomiting [15]. Therefore, de novo pyrimidine biosynthesis, and ODCase in particular, is a potential and challenging target for developing novel therapeutics. One must note that ODCase requires that the nucleoside ligands carry a monophosphate group at the 5´-position, which is responsible for a significant amount of binding energy for the initial binding

10.4155/FMC.13.198 © 2014 Future Science Ltd

Future Med. Chem. (2014) 6(2), 165–177

Sourabh Mundra1,2 & Lakshmi P Kotra*1,3 Center for Molecular Design & Preformulations, & Toronto General Research Institute, University Health Network, Toronto, Ontario, M5G 1L7, Canada 2 Birla Institute of Technology & Science, Pilani, Rajasthan, 333031, India 3 Department of Pharmaceutical Sciences, Leslie Dan Faculty of Pharmacy, University of Toronto, Ontario, M5S 3M2, Canada *Author for correspondence: Tel.: +1 416 581 7601 E-mail: [email protected] 1

ISSN 1756-8919

165

Review | Mundra & Kotra O O

O-

HN

O

PPi

PRPP

HN H3N

O-

L-aspartate

O

O O3PO

-2

+

O

N H Orotate

O

N

O

COOOrotate phosphoribosyl transferase

HN COO-

OH OH OMP (1)

H+

CO2 O3PO

-2

Orotidine monophosphate decarboxylase

N

O O

OH OH UMP (2)

Figure 1. De novo biosynthesis of uridine monophosphate from the amino acid, l-aspartate. Orotate phosphoribosyl transferase and OMP decarboxylase enzymatic functions are a part of one single enzyme, UMP synthase, in some species such as human, and others have these two enzymes as monofunctional enzymes. OMP: Orotidine monophosphate; UMP: Uridine monophosphate.

of the ligand to the catalytic site. However, such phosphorylated nucleosides, carrying a highly polar monophosphate group, would not be able to enter cells. Thus, for any biological activity in cell-based assays, one must either mask the polar groups or use the corresponding nucleoside forms and expect that cellular kinases would phosphorylate these nucleosides once they enter the cell. Due to this requirement, most of the cell assays have been carried out using the nucleoside forms of the inhibitors of ODCase, as is the typical practice for nucleoside drugs. This review will first discuss the mechanism of catalysis and the structure of the ODCase enzyme, leading into a discussion on multiple approaches to ODCase drug discovery, emphasizing the varied strategies that are likely to provide important new drugs in the future.

Key Terms Bifunctional enzyme:

Contiguous protein molecule carrying two catalytic sites with ability to conduct two distinct biochemical reactions.

De novo pathway: Synthesis of complex molecules from simple molecules in a cell.

Salvage pathway: Synthesis

of nucleotides from intermediates in the degradative pathway for nucleotides.

Direct decarboxylation:

Carboxylation of orotidine-5´monophosphate without involving a covalent intermediate, but induced by stress and a variety of other factors.

166

Catalytic mechanism of ODCase Many investigative studies were undertaken based on chemical and biochemical rationale, x-ray crystal structures and computer simulations, in an attempt to elucidate the mechanism of catalysis of ODCase. Here, we catalog the main mechanisms that were proposed over the past few decades, although since 2000, there has been an intense research into confirming the most possible mechanism of decarboxylation by ODCase based on the x-ray structures of ODCase, mutants of ODCase, and through probe molecules. Beak and Siegel, proposed the ‘nitrogen ylide’ mechanism as a result of examining the non-enzymatic decarboxylation of 1,3-dimethylorotic acid and related compounds. They proposed that the reaction is initiated by substrate (OMP) protonation at O4 to yield a betaine structure with a positive charge localized at N-1 (Figure 2A) [16]. The positive charge then, by inductively stabilizing the adjacent vinylic Future Med. Chem. (2014) 6(2)

carbanion of the intermediate state, facilitates the decarboxylation reaction yielding UMP. A second mechanism, proposed by Silverman and Groziak, involves a Michael addition at the C5 position of the pyrimidine ring followed by decarboxylation (Figure  2B) [17]. Nucleophilic attack by the active site occurs at the C5 position of OMP followed by enzyme-mediated proton donation. This causes a change in the geometry at C5 from a trigonal (sp2) to a tetrahedral (sp3) center, which facilitates the decarboxylation/elimination of the trans intermediate to release UMP as the product. The basis for this mechanism was the observation that orotic acid analogs are susceptible to nucleophilic attack at C5. This proposal was later ruled out when Acheson and co-workers demonstrated a lack of any secondary deuterium isotope effect with the C5 hydrogen [18]. A mechanism involving direct decarboxylation of OMP was proposed after the x-ray structure of ODCase was resolved (Figure 2C) [9], based on x-ray structure ana­lysis of various ODCase inhibitor complexes [19–22]. According to this mechanism, the decarboxylation is initiated by charge repulsion experienced by the carboxyl group on OMP and the Asp residues in the catalytic site of ODCase, followed by proton transfer from Lys73 to the C6 position of the pyrimidine moiety. The addition of the proton is able to destabilize the ground state and stabilize the transition state, which facilitates the decarboxylation of OMP. Computational models investigated by Warshel et al. suggest the ground-state destabilization further supporting this mechanism [23]. Gao and co-workers proposed that the proximity of the aspartate residue to OMP carboxylate, in the active site, produces a ‘ground-state destabilization’ effect that facilitates direct decarboxylation [24]. Recently, there has been more evidence future science group

Design of inhibitors of ODCase towards this mechanism, involving stress onto the carboxyl group at the C6 position of OMP and elimination of the CO2 molecule or direct decarboxylation facilitated through the stress [25]. Lee and Houk proposed a mechanism with a carbene intermediate for the enzymatic decarboxylation [26]. They based their mechanism on quantum mechanical calculations for the ground state and the energies of potential intermediates during OMP decarboxylation. They proposed that protonation at O4 gave a further stabilized intermediate that has both a zwitterionic and a neutral carbene resonance structure. In contrast to Beak and Siegel, this proposal predicts that a positive charge at N1 in OMP would not be important in the enzymatic reaction. This proposal could not be supported by any experimental means and is not widely accepted. Another proposed catalytic mechanism (F igure  2D) by Kollman assumes that Asp70 – one of the two Asp residues in the catalytic tetrad – is protonated [27], and predicts that the decarboxylation is initiated by an enamine protonation at the C5 position by Lys72 (according to the Methanobacterium thermoautotrophicum (Mt) ODCase numbering). Hur and Bruice, studying molecular dynamics simulations of the ground state and intermediate/transition states of the ODCase/OMP structure, provided additional support for the idea of stabilization of the O

O

N

+H + N

HO

O +H

HN O

N

Nu

O

O

-CO2 O

HN

+H N R I O

O

HN O

N R II

N R IV

N

O HN

+H

O

O

R III

O

R II

I O

HN

HN O

N

O

+ N

O

HN

O

O

HO

O

N

-CO2

HN

R II O

-CO2

O

R III

H

N

O

HN

R

Nu

HN

O

O

O

R II O

O

R I

O

HN

+H O

R I O

O

carboanionic intermediate by a nearby lysine and reorganization of the 203–218 loop [28]. In addition, they found that the intermediate and the transition state have similar associated energies, thus, the stabilization of the former reduces the free energy of the latter. Shostack and Jones proposed the Schiff base formation mechanism in which the active site lysine residue plays an important role [29]. However, the lack of 18O exchange of O4 with water seems to rule this out as a possible mechanism. Appleby and co-workers proposed a concerted mechanism for the decarboxylation. In this mechanism, C6 of OMP is protonated while the carboxylate bond is broken [30]. Accordingly, C6 would start forming a bond with hydrogen on the catalytic Lys residue in the active site, as the p-orbital of C6 turned to the e-ammonium group of the same catalytic lysine, creating the transition state. The carboxylate group is still loosely linked to the C6 while a gradual elongation of the C6–C7 bond occurs, facilitating the decarboxylation. This could explain the experimental results reported by Radzicka and Wolfenden [3]. However, recent studies have confirmed that concerted mechanism for decarboxylation can be ruled out, in favor of direct decarboxylation (vide infra) [31,32]. Jordan and Patel conducted a comparative study of three types of decarboxylases in the

O

HN

| Review

N R III

H

O H

-CO2

H

HN O

N R

O H

HN O

III

N R

IV

Figure 2. Proposed mechanisms of decarboxylation by ODCase. (A) Nitrogen-ylide mechanism, (B) nucleophilic addition mechanism, (C) direct decarboxylation, (D) C5-protonation mechanism.

future science group

www.future-science.com

167

Review | Mundra & Kotra

Ligand binding site

3.2 Å

Asp141

3.4 Å 3.3 Å

2.9 Å Lys138

Lys102 Asp136

Figure 3. 3D structure of ODCase from Plasmodium falciparum (PDB code: 2ZA2). Catalytic tetrad is shown in a capped stick representation and the binding site is shown as a Connolly surface. Secondary structure of the ODCase backbone is shown in a cartoon representation. Hydrogen bonds between the functional groups in the catalytic tetrad are shown in broken lines in orange. Residue numbering is as per the P. falciparum ODCase sequence.

context of their catalytic mechanisms of decarboxylation, including ODCase, and highlighted the challenges in explaining the unusual catalytic mechanism of ODCase. While there are common features between decarboxylases such as pyridoxal phosphate-dependent decarboxylases and ODCase, there are some differential features with respect to substrate binding [5]. Using computational approaches, that is, hybrid quantum-molecular mechanics simulations, decarboxylation of OMP and a close analog, 5-fluoro-OMP were conducted. An integral role for Lys72, one of the two lysine residues in the catalytic tetrad, in stabilizing the transition state during the decarboxylation of OMP and the direct decarboxylation mechanism were confirmed [33]. Fujihashi et al. have recently determined a 1 Å-resolution structure of ODCase, including a Lys72 mutant, in complex with the product UMP. Their results strongly support the notion that electrostatic stress and substrate distortion contribute to the direct decarboxylation of OMP [25]. A relatively non-polar microenvironment within the active site in turn may also enhance the strength of electrostatic bonds. One must note that there are other effects such as 168

Future Med. Chem. (2014) 6(2)

strain on the pyrimidine ring, indirect effect of the role of 5´-monophosphate for the binding of the ligand and so on, which ultimately lead to the disruption of the electronic conjugation of the pyrimidine ring. Thus, the contributions of a number of these factors collectively explain the very high catalytic efficiency of this enzyme, which cannot be explained by any single factor. In a separate set of experiments, Gerlt, Richard and co-workers conducted ODCase-catalyzed deuterium exchange reactions between solvent D2O and 5-fluoro-UMP, UMP and the truncated fluorinated substrate FEU [34–37]. These experiments provide strong support for direct loss of CO2 molecule through a vinyl carbanion reaction intermediate. Structures of ODCase The function and enzymatic activity are intrinsically linked to the structure of the ODCase enzyme. To date, 152 ODCase x-ray crystal structures from both prokaryotic and eukaryotic organisms have been solved and appear in the Protein Data Bank [201]. This large collection of ODCase structures along with various mechanistic studies have led to a slow but steady increase of our understanding of the underlying biochemical mechanism for its catalytic activity. As described above, there have been many mechanisms proposed for the decarboxylation of OMP to UMP, and some uncertainty may still exist on the catalytic mechanism. x-ray crystallography of ODCase in complex with its various ligands helps in understanding the interaction between the ligands with the amino acids of the active site and as a result, the active site of this enzyme has been well mapped [22,38,39]. The catalytic site of ODCase consists of two aspartate residues (Asp136 and Asp141B, the latter contributed by the second subunit of the dimeric ODCase; Plasmodium falciparum (Pf ) ODCase numbering) and two lysine residues (Lys102 and Lys138) that are held in place by a strong network of hydrogen bonds (Figure 3) [22]. Analyses of several co-crystal structures of ODCase with a variety of ligands confirm that these residues are in close proximity to substitutions at the C6 position of the pyrimidine in various nucleotide ligands, including the product UMP [22]. The ribosyl moiety and the 5´-monophosphate are involved in an extensive hydrogen bonding network and the resulting binding energy between the enzyme and the ligands is the driving force to place the pyrimidine base into the active site pocket [22,25,39–41]. A recent study has shown that future science group

Design of inhibitors of ODCase purine nucleotides, in addition to pyrimidines, also inhibit ODCase, such as xanthosine monophosphate (XMP; 28), GMP, AMP and IMP [22,42]. Interestingly ODCases from diverse species exhibit remarkable differences in their affinity to various ligands. For example, XMP binds approximately 150-fold tighter to the plasmodial ODCase than to the human enzyme [39,42]. ODCase continues to attract attention due to its unusual catalytic efficiency, the discovery of a variety of ligands its active site can accommodate and the new twists it imparts onto the ligands. Along these lines, two cytidine-based analogs undergo quite an interesting rearrangement facilitated by ODCase within the catalytic site, providing ideas for novel inhibitors design [43]. Comparison of known structures of ODCase from various sources, namely bacteria, the malaria parasite and humans will illuminate the conserved residues in catalysis of this most proficient enzyme. Considering the vast repertoire of this structural information, this could be the subject of a separate article. However, this is important knowledge to explore the design of inhibitors targeting ODCase when a structure-based approach is undertaken for drug design. Nucleoside inhibitors of ODCase as anticancer agents Nucleosides and compounds bearing nucleosidelike features interfering with the pyrimidine biosynthesis pathway have shown interesting anticancer activities [44]. 5-fluorouracil and methotrexate are two such examples of drugs that are able to impair nucleotide synthesis in cancer cells, and thereby functionally reduce the uncontrolled proliferation of these cells [45]. During uncontrolled and fast replication of cells, the de novo pathway must also contribute toward supplying nucleotides despite the existence of the salvage pathway because of the higher demand for nucleotides in a short period of time [46]. 6-aza-uridine (17) and pyrazofurin (3) exhibited good anticancer activities against a number of clinical tumor models [47,48]. Pyrazofurin (3) is a C-nucleoside antibiotic that also blocks de novo pyrimidine synthesis by inhibiting ODCase. In vivo, it is converted to its 5´-phosphate derivative (4) and then engages in competitive inhibition of the ODCase enzyme (UMP synthase). The 5´-monophosphate derivative of pyrazofurin (4) was originally isolated from the broth filtrates of Streptomyces candidus [101] and is a potent inhibitor of ODCase with inhibition constants in the low nanomolar range future science group

| Review

(Figure 4) [49].

It has a breadth of clinically relevant biological activities and subsequently has been the focus of numerous investigations as an anticancer agent [50]. Pyrazofurin was shown to inhibit the growth of several solid animal tumors but has no effect on most transplantable carcinomas [51]. Pyrazofurin was considered a breakthrough molecule in late 1970s and early 1980s, and was considered for clinical development. However, pyrazofurin (3) had a limited antitumor effect with toxicity mostly to the oral mucosa in Phase I trials and was not pursued any further [52,53]. An x-ray crystal structure of the complex of pyrazofurin-5´-monophosphate (4) and human ODCase indicated a very strong network of hydrogen bonds between the ligand and the enzyme [10]. Interestingly, 4 is a competitive inhibitor of Mt ODCase (K i = 6.2 ± 0.6 µM), O H2NOC

RO

HN

NH N

HO O

HO

O

R

HO

O

N NH

H2NOC

OH

8 R = COOMe 9 R = CONH2

HN N

NH

O

HO

CONH2

CONH2

O

OH OH

OH OH

N N H

7

5 R=H 6 R = PO3 2-

CONH2

OH

OH OH

OH OH

3 R=H 4 R = PO32-

HO

I

O

OH OH

S

N

O RO

OH

10

OH

11

O NH

N O HO

N

N

NH2

O Br

OH

N H

OH

12

H N N

N

O

NH2

O

HO

CONH2

14

13

NO2 NO2 COOMe

MeOOC N H

15 (Nifedipine)

O

O

O

O

OMe

N H 16 (Nimodipine)

Figure 4. Structures of nucleoside-like inhibitors of ODCase and non-nucleoside ODCase inhibitors. (A) Nucleoside-like inhibitors of ODCase, (B) non-nucleoside ODCase inhibitors.

www.future-science.com

169

Review | Mundra & Kotra Table 1. C6 substituted and C5 fluorinated uridine monophosphates that exhibit potent ODCase inhibition and anticancer activities. Compound number

Compound

18

6-aza-uridine-5´-Omonophosphate 5-fluoro-6-azidouridine-5´-Omonophosphate 5-fluoro-6-aminouridine-5´-Omonophosphate 5-fluoro-6-ethyluridine-5´-Omonophosphate 6-azidouridine -5’-O-monophosphate 6-aminouridine -5’-O-monophosphate

19 20 21 23 24

Mt ODCase Ki (µM) Mt ODCase

Hs ODCase

11

NA

NA

0.36 ± 0.03

16.60 ± 0.70

11.40 ± 0.60

0.35 ± 0.01

29.00 ± 0.30

NA

0.20 ± 0.10

NA

0.84 ± 0.02

Enzyme kinetics conducted in different laboratories may have used distinct experimental conditions, and comparisons must be made with care. Hs: Homo sapiens; Mt: Methanobacterium thermoautotrophicum; NA: Not available. Reproduced with permission from [47] © American Chemical Society (2009).

but inhibits human ODCase, as a slow tightbinding inhibitor with nanomolar potency (estimated K i = 17 nM) [10]. 6-aza-uridine (17) is also converted into its mononucleotide form in vivo, 6-aza-uridine-5´-O-monophosphate (18), and the resulting mononucleotide inhibits ODCase, impairing the de novo production of pyrimidine nucleotides [54]. This compound inhibits yeast and Mt ODCases with the inhibition constants (K i) of 64 nM and 11 µM, respectively (Table 1) [55,56]. Kotra and co-workers have developed a variety of C6 substituted 5-fluoro-UMPs derivatives that exhibit potent ODCase inhibition and evaluated their anticancer activities (Table 1) [46]. These compounds carried small substitutions at the C6 position of the 5-fluoropyrimidinyl backbone, including cyano, azido, amino, iodo, N-methylamino, N,N-dimethylamino, aldehyde, ethyl ester and hydroxyl amino moieties. The inhibition of Mt ODCase by various inhibitors in Table 1 were evaluated in a competitive inhibition assay and the reversible inhibition of human ODCase was studied at 37°C using isothermal titration calorimetry, a novel method developed by Kotra and co-workers [56]. The reaction rates were measured by monitoring the heat generated during the decarboxylation reaction of orotidine5´-monophosphate. The presence of the fluorine moiety at C5 of uridine or UMP was expected to offer a stabilizing effect to the C6 substitutions as well as generate an electrophilic center at C6 due to the electronegativity of the fluorine moiety [46]. These mononucleotide derivatives were evaluated for ODCase inhibitory activities, 170

Future Med. Chem. (2014) 6(2)

while the corresponding nucleoside derivatives were evaluated in the cellular assays. The parent compound, 5-fluoro-UMP was a very weak inhibitor of ODCase, but 6-azido-5-fluoro (19) and 5-fluoro-6-iodo derivatives were covalent inhibitors of ODCase, akin to their structural analogs without the fluorine moiety. The active site residue Lys145 (human ODCase) covalently binds to the ligand after the elimination of the 6-substitution, which is clearly seen in the cocrystal structure of the ligand–enzyme complex, supporting the enzyme kinetics [46]. Among the synthesized nucleoside derivatives, 6-azido-5-fluoro, 6-amino-5-fluoro and 6-carbaldehyde-5-fluoro (22) derivatives showed potent anticancer activities in cell-based assays against various leukemia cell lines. Specifically, 6-azido, 6-amino, 6-ethyl and 6-aldehyde derivatives of 5-fluorouridine exhibited potent anticancer activities in cell lines of hematopoietic origin, including leukemias (acute myelogenous leukemia lines OCIAML-1 and OCI-AML-2, and T-cell leukemia line, SKW3), lymphomas (non-Hodgkin’s B cell lymphoma-OCI-Ly-7), and multiple myeloma (OCI-My-2), with IC50 values in the range of 0.3–2.1 µM. Adherent cancer cell lines, including breast cancer lines MCF7 and MDA468, were generally less sensitive to these derivatives, with IC50 values ranging between 1 and 31 µM [46]. There appears to be some selectivity towards cancer lines when compared with the cytotoxicity of these compounds to concanavalin A-activated healthy human PBMC cells, offering the possibility that ODCase inhibitors could have potential as anticancer agents. ODCase (i.e., UMP synthase in humans) as an anticancer target will need to be further explored, but there is a prospect to design inhibitors to this enzyme to inhibit cancer cells due to its importance in pyrimidine biosynthesis. Potent inhibitors of de novo pyrimidine biosynthesis may have limited value as anticancer drugs because of the important cellular roles of activated intermediates such as UDP-glucose and CDP-choline. Consistent with this, PALA and pyrazofurin are clinically too toxic to be used as anticancer drugs. Nucleoside inhibitors of ODCase as antivirals Viral infections are unique since the invading virus becomes a component of the host cell, integrating with and utilizing the host cell’s biosynthetic machinery. In this aspect, the chemical modification of nucleosides has become one future science group

Design of inhibitors of ODCase of the most important chemical strategies to design antiviral agents. However some of these agents have targeted host enzymes rather than viral-specific enzymes. ODCase thus has been a potential target for drugs directed against RNA viruses such as pox and flaviviruses, and potentially for antibacterial drugs [57–59]. Pyrazofurin (3) was evaluated as an antiviral agent for respiratory syncytial virus, influenza, vaccinia, West Nile virus and other RNA viruses [60–63]. Pyrazofurin (3) and its a-epimer, pyrazofurin B (7) exhibit a broad spectrum of antiviral and antitumor activities at low concentrations, but also show significant toxicity [59–66]. Reduced toxicity was observed by replacing the pyrazole ring in pyrazofurin with a thiophene ring [64] in such compounds as methyl-5-carboxamido4-hydroxy-3-(b-d-ribofuranosyl)-2-thiophene carboxylate (8) and its corresponding 2,5-thiophene dicarboxamide derivative (9), two new analogs of pyrazofurin (Figure 4). Certain nucleoside analogs have been found to exhibit an ability to stimulate the immune system, which is important for fighting viruses. For example, 8-substituted (8-bromo and 8-mercapto) guanosines and phenyl-substituted pyrimidinones possessing a guanine-like functionality can activate B cells and natural killer cells, induce interferons and exhibit antiviral activity in vivo [67–74]. Another related compound, 7-methyl-8-oxoguanosine (12), is also reported to be a B-cell activator [75]. In these series of compounds, the important molecular configurations for this activity appear to reside in the pyrimidine portion of the molecule (resembling guanosine) and in the substitutions at the 7 and 8 positions on the guanine moiety. Wicker et al. have suggested the role of a biochemical pathway common to several immune cell types in immune cell activation [68]. Interestingly, adenosine- or inosine-like structures do not appear to be immuno potentiators. 2-amino-5-bromo6-phenyl-4-pyrimidinone or bropirimine (13) is one such compound that is extensively studied and exhibits antiviral activity. The pyrazolebased nucleoside 3-(2-hydroxyethoxymethyl] pyrazole-5-carboxamide (14), which is an analog of 4-deoxypyrazofurin-3-(b-d-ribofuranosyl)pyrazole-4-carboxamide (10), also possesses antiviral activity (Figure 4) [71]. Nucleoside inhibitors of ODCase as antimalarials Malaria is caused by Plasmodia parasites, specifically Pf, Plasmodium vivax and Plasmodium future science group

| Review

ovalae, and increasingly Plasmodium malariae and Plasmodium knowlesi as the causative agents for human malaria cases [76,77]. The de novo pathway for the biosynthesis of pyrimidine nucleotides in Plasmodia provides essential precursors for the synthesis of DNA and RNA. Unlike the cells of its human host, Plasmodia are unable to salvage preformed pyrimidines from the extra­cellular environment and therefore, are restricted to using the de novo pathway for pyrimidine nucleotide synthesis, the only one operational inside the parasite cell [78,79]. Since mature human red blood cells lack the de novo pyrimidine pathway [80], inhibiting this pathway could selectively kill the parasites that infect these red blood cells. The de novo pathway includes six enzyme-catalyzed reactions, the last of these being the ODCase catalyzed decarboxylation to produce UMP. ODCase in fact is a very attractive target for designing antimalarial agents due to the essential nature of this enzyme for Plasmodia [81–83]. A number of C6 substituted UMPs were synthesized using the principles of bioisosterism (Table 2). A structure–activity relationship study with strategically placed groups at the C6 position of uridine and UMP, such as 6-iodo, 6-cyano, 6-azido, 6-amino, 6-methyl, 6-N-methyl and 6-N,N-dimethyl groups revealed interesting findings about the binding of these nucleotides at the Table 2. Inhibition kinetics of C6-substitued uridine monophosphate derivatives and xanthosine monophosphate against Plasmodium falciparum ODCase. Compound number Compound

Pf ODCase Ki (µM)

2

UMP 6-aza-UMP Allopurinol-MP Pyrazofurin-5´-O monophosphate

210 ± 10 0.012 ± 0.003 0.24 ± 0.02 0.0036 ± 0.0007

[44]

6-azido-UMP 6-amino-UMP 6-methyl-UMP 6-cyano-UMP Xanthosine 5´-O monophosphate 6-iodo-UMP CMP-N3-oxide 6-NHMe-uridine -5’-O-monophosphate 6-N(Me)2-uridine†

2.0 ± 0.1 2.1 34.1 ± 0.4 26 ± 2 0.0044 ± 0.0007

[87]

18 25 4 23 24 26 27 28 6 29 30 31

Ref. [39] [39] [39]

[87] [87] [44,87] [39] [39]

6.2 ± 0.7 22.1 ± 3.2 28.5 ± 8.9

[44]

31.7 ± 10.0

[87]

[87]

Enzyme kinetics conducted in different laboratories may have used distinct experimental conditions, and comparisons must be made with care. † This molecule is tested in its 5’-O-monophosphate form against Pf ODCase. Pf: Plasmodium falciparum; UMP: Uridine monophosphate.

www.future-science.com

171

Review | Mundra & Kotra Key Term Promiscuous: Ability of

ODCase (or any other enzyme) to facilitate biochemical reactions other than its native decarboxylation, and also its ability to bind to molecules with a variety of structures as ligands.

catalytic site of ODCase, and the antimalarial activities of the corresponding nucleoside analogs in Pf cultures [84,85,201]. 6-iodo-UMP (6) and 6-azido-UMP (23) covalently inhibit ODCase, while 6-cyano-UMP (27) and 6-amino UMP (24) are competitive inhibitors [42,86,87]. 6-aza-UMP (18) is a moderate inhibitor of Mt ODCase with a K i of 12.4 ± 0.7 µM, but it inhibits Pf ODCase with higher potency (K i = 1.1 ± 0.03 µM) (F igure  5) . 6-cyano-UMP (27), an apparent competitive, non-covalent inhibitor, inactivates Pf ODCase with a K i of 29 ± 2 µM (Mt ODCase, 26 ± 2 µM), but against P. vivax ODCase it exhibits a weaker inhibition with a K i of 62 ± 5 µM [57,85,88]. However, in time-dependent studies, upon incubation of 6-cyano-UMP (27) with Pf ODCase for several hours, barbiturate-5’-monophosphate (BMP) is generated and this transformation also occurs with Mt ODCase [86,88]. The inhibition constant (Ki) for 6-methyl-UMP (26) is O

O

HN N

RO

N

O3PO

OH

O

17 R = H 18 R = -PO32-

O3PO

N

OH

OH

O3PO

2-

O

2-

O3PO

27

28

OH OH 25 O

NH N H

O

NH2 N

O

O3PO

2-

N

O OH

OH

29

O

HN

HN N

NHCH3

O HO

O OH

O OH OH

OH OH

O

O3PO

N

CN

N

O

O N

O

26

O

O3PO

24

N

N

N 2-

OH OH

OH

O

NH2

O

HN N

HO N

O

OH OH

2-

O3PO

23 HN

O OH OH 21

N

O 2-

O

O3PO

N3

O

22

2-

OH

C2H5

N

HN

O 2-

O

O

O3PO

2-

O

HN

CHO

NH2

O

O

N

O

N

OH

F

HN

20

F

HN

OH

O3PO

19

O

HO

O 2-

OH OH

OH

O

N3

N

O

F

HN

O 2-

O

O

F

HN

O

a competitive inhibitor of ODCases from Pf and P. vivax with moderate potency (K i = 34.1 ± 0.4 and 22.4 ± 0.7 µM, respectively). 6-amino-UMP (24) inhibited Mt ODCase competitively with a K i of 840 ± 25 nM, but showed slightly less potency against Pf ODCase (K i ~2.1 ± 0.0 µM) [85]. When evaluated as a noncovalent inhibitor, 6-azido-UMP (23) is a good inhibitor of Mt and Pf ODCases (K i = 0.19 ± 0.01 and 2.0 ± 0.1 µM, respectively). However, in time-dependent assays 6-azido-UMP (23) inactivated Mt and Pf ODCases with first-order rates of inactivation (kinact) of 61.2 and 0.35 h-1, respectively. By comparison, 6-iodo-UMP (6) inhibited Pf ODCase with a second-order rate constant (k obs/[I]) of 680,000 M-1sec-1; 6-azido-UMP (23) inhibited this enzyme with the second-order rate constant (kinact/K I) of 13.2 M-1sec-1, clearly indicating the superior inactivation of ODCase by 6-iodo-UMP (6) [85,88]. In certain cases, kinact /K I could not be

OH

N

N(CH3)2

O OH

30

OH

31

Figure 5. Various nucleoside inhibitors of ODCase with antimalarial and anticancer activities.

172

Future Med. Chem. (2014) 6(2)

future science group

Design of inhibitors of ODCase derived and the second-order rate constant is estimated to be kobs /[I], and were used for comparative purposes only. Derivation of these parameters is pertinent to the type of irreversible inactivation that a compound exhibits. kobs /[I] (slope of a linear fit) indicates one-step inactivation without the reversible enzyme–inhibitor complex formation, where as kinact /K I is derived from a non-linear (hyperbolic) profile indicating the presence of reversible complex formation before the irreversible covalent bond is formed. This discussion is elaborately included in the reference [46] in the context of ODCase inhibitors, and a general discussion of the enzymology can be gleaned from reference [88]. Several 6-substituted uridine derivatives were evaluated in cell-based assays as potential antimalarial agents. 6-N-methylamino uridine and 6-N,N-dimethylamino uridine (31) were moderate inhibitors of Pf with IC50 values of 28 ± 9 and 32 ± 10 µM, respectively. 6-methyluridine exhibited very weak activity against the Pf 3D7 isolate (IC50 of 530.5 ± 0.1 µM) although its 5´-monosphosphate derivative (26) is a potent inactivator of ODCase. 6-methyluridine did not show toxicity to CHO cells (IC50 > 1.2 mM). 6-N-methylamino uridine and 6-N,N-dimethylamino uridine (31) inhibited CHO cells with IC50 values in the range of 68–72 µM (Figure 5). Interestingly, 6-iodouridine (5), the most potent inhibitor of Pf ODCase, other than BMP, exhibited good antimalarial activity against several strains of Pf ranging from 1.2 ± 0.2 to 56.5 ± 14.0 µM including mefloquine-resistant and chloroquine-resistant isolates [84,90]. 6-iodouridine (5) inhibited drug-resistant isolates ItG, and KC98 at low micromolar concentrations, while multidrug resistant isolates such as MB and SB were less susceptible to 6-iodouridine (5). Interestingly, the human parasite Pf exhibited much higher sensitivity, almost 25-fold higher, than the rodent parasites Plasmodium berghei and Plasmodium chabaudi chabaudi, to 6-iodouridine (5). This compound did not generate rapid drugresistant populations of the parasite in vitro [89]. 6-iodouridine (5) was evaluated in combination with artemisinin and azithromycin, in malaria mouse models, and showed good efficacy in vivo with artemisinin and azithromycin. 6-iodouridine exhibited an additive effect in combination with artemisinin in vitro, and in an in vivo study; a drop in parasitemia was observed during treatment but a sharp ‘rebound’ in the parasitemia was observed when the drug treatment was stopped. Overall, however, the combination of 6-iodouridine and artemisinin showed a beneficial outcome in this future science group

| Review

study [89]. A combination of 6-iodouridine with azithromycin indicated that the combination is significantly better than either treatment alone and there was no ‘rebound effect’ observed due to the longer half-life of azithromycin than that for artemisinin [89]. Interestingly, supplementation of uridine when treated with 6-iodouridine reduced any drug toxicities to the mouse, thereby facilitating the administration of higher doses of the drug to achieve higher efficacy. This validates the mechanism of these ODCase inhibitors that pyrimidine nucleosides can be salvaged in the host cell but not by the malaria parasite [89], which is an important observation useful for the development of this molecule as a potential antimalarial agent. Investigations into the utility of ODCase inhibitors targeting other parasitic infections are still at an early stage. During a recent study using Trypanosoma brucei ODCase, it was shown that this enzyme may not be the primary target for pyrazofurin-5´-monophosphate (4) and the organism can still survive when UMP synthase is not expressed [90]. However, the parasite virulence is lost when a double-knockout mutant of UMP synthase from T. brucei was generated, and this mutant regained virulence after prolonged culture implying its ability to survive and evolve using the salvage pathway for obtaining UMP through the host cellular machinery [90]. Non-nucleoside inhibitors for ODCase Two non-nucleoside inhibitors, nifedipine (15) and nimodipine (16; Figure 4B), clinically used as calcium channel blockers, have been reported to inhibit ODCase competitively, with inhibition constants (K i) of 105 and 18 µm, respectively. Although ODCase may have evolved to carry out the decarboxylation of OMP with high efficiency, it appears to be an inherently promiscuous enzyme, binding to other nucleosides, both purines and pyrimidines, as well as to non-nucleoside inhibitors [10,91]. Based on the crystal structure of the complexes of various ODCase inhibitors such as BMP, pyrazofurin-5´-monophosphate (4), 6-cyanoUMP (27), 6-aza-UMP (18) as well as non-nucleoside inhibitors, Meza-Avina et al. proposed an empirical pharmacophore describing the characteristics of potent inhibitors to ODCase [10]. According to this model, three separate regional characteristics were identified on the pharmacophore that various compounds should possess as inhibitors of ODCase. These general principles help simplify the complex network of interactions between ODCase and its various ligands. www.future-science.com

173

Review | Mundra & Kotra Future perspective There is strong support that ODCase is a good drug target and a potentially druggable target to design potent inhibitors. With the availability of a large set of x-ray crystal structures, many in complex with ligands, ODCase provides an opportunity for new drug discovery. Due to its essential role in pyrimidine biosynthesis in a variety of organisms’ genomes, such inhibitors could be developed for a variety of clinical indications. ODCase inhibitors have shown promise as potential drug candidates against Plasmodia parasites, although this mechanism could also be applicable in other parasitic diseases and remains to be explored. Development of ODCase inhibitors as anticancer agents may have some merit, especially against fast-growing cancers because cancer cells would depend on the de novo supply of pyrimidines for their fast division. It is important to consider the fundamental nucleotide biosynthesis and the role of inhibition of ODCase in affecting the targeted

disease condition for any clinical development, because nucleotide biosynthesis and its integrity have evolved differently across various species. Acknowledgement The authors are grateful to EF Pai for discussions on the mechanisms of decarboxylation by ODCase. The authors acknowledge W Wei for the preparation of Figure 3.

Financial & competing interests disclosure The authors wish to thank ISTP Canada (LPK; Grant #ICRD08–15) and the Department of Biotechnology, Ministry of Science and Technology, Government of India for financial support. LP Kotra has financial interest in the company that licensed the disclosures in patent number US8067391. The authors have no other relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript apart from those disclosed. No writing assistance was utilized in the production of this manuscript.

Executive summary Background ODCase is an important component of the de novo nucleic acid biosynthesis pathway. Catalytic mechanism „„

There is strong evidence that ODCase decarboxylates orotidine monophosphate via direct decarboxylation imparting stress onto the substrate, that is, the C6 carbanion-based transition-state intermediate appears during the early stage of the reaction, and a variety of stress, including electrostatic, appears to play a role in the process of decarboxylation of the substrate. Nucleoside inhibitors „„

C6-subsituted derivatives such as 6-azido-uridine monophosphate and 6-iodo-uridine monophosphate are covalent inhibitors of ODCase. Structures of ODCase „„

A plethora of 3D structures of ODCase and numerous studies on the inhibition mechanisms provide ample opportunity to design novel inhibitors to ODCase. Nucleoside inhibitors as anticancer agents „„

C6-substituted 5-fluorouridine derivatives and other similar derivatives may possess useful anticancer activities, targeting ODCase. Nucleoside inhibitors of ODCase as antimalarials „„

„„

A number of C6-substituted uridine monophosphate derivatives with small substitutions exhibit potent antiplasmodial activities. 6-iodouridine is a potential antimalarial agent with a novel mechanism of action.

References Papers of special note have been highlighted as: n of interest nn of considerable interest 1

Reichard P. The enzymatic synthesis of pyrimidines. Adv. Enzymol. Mol. Biol. 21, 263–294 (1959).

2

Callahan BP, Miller BG. OMP decarboxylaseAn enigma persists. Bioorg. Chem. 35(6), 465–469 (2007).

3

Radzicka A, Wolfenden R. A proficient enzyme. Science 267 (5194), 90–93 (1995).

174

nn

4

5

Describes ODCase catalytic proficiency.

6

Lewis CA Jr, Wolfenden R. Uroporphyrinogen decarboxylase as a benchmark for the catalytic proficiency of enzymes. Proc. Natl. Acad. Sci. USA 105(45), 17328–17333 (2008).

Donovan WP, Kushner SR. Purification and characterization of orotidine-5´monophosphate decarboxylase from Escherichia coli K-12. J. Bacteriol. 156 (2), 620–624 (1983).

7

Pragobpol S, Gero AM, Lee CS, O’Sullivan WJ. Orotate phosphoribosyltransferase and orotidylate decarboxylase from Crithidia luciliae: subcellular location of the enzymes and a study of substrate channeling. Arch. Biochem. Biophys. 230(1), 285–293 (1984).

Jordan F, Patel H. Catalysis in enzymatic decarboxylations: comparison of selected cofactor-dependent and cofactor-independent examples. ACS Catal. 3(7), 1601–1617 (2013).

Future Med. Chem. (2014) 6(2)

future science group

Design of inhibitors of ODCase 8

Warshel A, Florian J. Computer simulations of enzyme catalysis: finding out what has been optimized by evolution. Proc. Natl Acad. Sci. USA 95(11), 5950–5955 (1998).

9

Miller BG, Wolfenden R. Catalytic proficiency: the unusual case of OMP decarboxylase. Annu. Rev. Biochem. 71, 847–885 (2002).

5´-monophosphate decarboxylase by crystallography. Biochemistry 41(12), 4002–4011 (2002). nn

20 Feng WY, Austin TJ, Chew F, Gronert S, Wu

W. The mechanism of orotidine 5´-monophosphate decarboxylase: catalysis by destabilization of the substrate. Biochemistry 39(7), 1778–1783 (2000).

10 Meza-Avina ME, Wei L, Liu Y et al.

Structural determinants for the inhibitory ligands of orotidine-5´-monophosphate decarboxylase. Bioorg. Med. Chem. 18(11), 4032–4041 (2010).

21 Begley TP, Appleby TC, Ealick SE. The

structural basis for the remarkable catalytic proficiency of orotidine 5´-monophosphate decarboxylase. Curr. Opin. Struct. Biol. 10(6), 711–718 (2000).

11 Gero AM, O’Sullivan WJ. Purines and

pyrimidines in malarial parasites. Blood Cells 16(2–3), 467–484 (1990). 12 Reyes P, Guganig ME. Pyrimidine

phosphoribosyltransferase from murine leukemia P1534J. Partial purification, substrate specificity, and evidence for its existence as a bifunctional complex with orotidine 5-phosphate decarboxylase. J. Biol. Chem. 250(13), 5097–5108 (1975).

22 Wu N, Pai EF. Crystal structures of inhibitor

complexes reveal an alternate binding mode in orotidine-5´-monophosphate decarboxylase J. Biol. Chem. 277(31), 28080–28087 (2002). 23 Warshel A, Strajbl M, Villa J, Florian J.

Remarkable rate enhancement of orotidine 5´-monophosphate decarboxylase is due to transition-state stabilization rather than to ground-state destabilization. Biochemistry 39(48), 14728–14738 (2000).

13 McClard RW, Black MJ, Livingstone LR,

Jones ME. Isolation and initial characterization of the single polypeptide that synthesizes uridine 5´-monophosphate from orotate in Ehrlich ascites carcinoma. Purification by tandem affinity chromatography of uridine5´-monophosphate synthase. Biochemistry 19(20), 4699–4706 (1980).

n

stress in catalysis: structure and mechanism of the enzyme orotidine monophosphate decarboxylase. Proc. Natl Acad. Sci. USA 97(5), 2017–2022 (2000). n

15 Casper ES, Vale K, Williams LJ, Martin DS,

Young CW. Phase I and clinical pharmacological evaluation of biochemical modulation of 5-fluorouracil with N-(phosphonacetyl)-l-aspartic acid. Cancer Res. 43(5), 2324–9 (1983).

17 Silverman RB, Groziak MP. Model chemistry

for a covalent mechanism of Action of orotidine 5´-phosphate decarboxylase. J. Am. Chem. Soc. 104(23), 6434–6439 (1982). 18 Acheson SA, Bell JB, Jones ME, Wolfenden

R. Orotidine-5´-monophosphate decarboxylase catalysis: kinetic isotope effects and the state of hybridization of a bound transition-state analogue. Biochemistry 29(13), 3198–3202 (1990). 19 Wu N, Gillon W, Pai EF. Mapping the active

site-ligand interactions of orotidine

future science group

Discusses the most probable mechanism based on electrostatic stress to be responsible for decarboxylation by ODCase.

25 Fujihashi M, Mito K, Pai EF, Miki K. Atomic

resolution structure of the orotidine 5´-monophosphate decarboxylase product complex combined with surface plasmon resonance ana­lysis: implications for the catalytic mechanism. J. Biol. Chem. 288(13), 9011–9016 (2013).

16 Beak P, Siegel B. Mechanism of

decarboxylation of 1,3-dimethylorotic acid. A model for orotidine 5´-phosphate decarboxylase. J. Am. Chem. Soc. 98(12), 3601–3606 (1976).

Describes one of the most probable mechanisms attributed to decarboxylation by ODCase.

24 Wu N, Mo Y, Gao J, Pai EF. Electrostatic

14 Meza-Avina ME, Wei L, Buhendwa MG,

et al. Inhibition of orotidine 5´-monophosphate decarboxylase and its therapeutic potential. Mini-Rev. Med. Chem. 8(3), 239–247(2008).

A comprehensive description of the interactions of various ligands in the binding site of ODCase.

nn

Atomic resolution x-ray crystal structure of ODCase.

26 Lee JK, Houk KN. A proficient enzyme

revisited: the predicted mechanism for orotidine monophosphate decarboxylase. Science 276(5314), 942–945 (1997). 27 Lee TS, Chong LT, Chodera JD, Kollman P.

An alternative explanation for the catalytic proficiency of orotidine 5´-phosphate decarboxylase. J. Am. Chem. Soc. 123(51), 12837–12848 (2001). 28 Hur S, Bruice TC. Molecular dynamic study

of orotidine-5´-monophosphate decarboxylase

www.future-science.com

| Review

in ground state and in intermediate state: a role of the 203–218 loop dynamics. Proc. Natl Acad. Sci. USA 99(15), 9668 -9673 (2002). 29 Shostak K, Jones ME. Orotidylate

decarboxylase: insights into the catalytic mechanism from substrate specificity studies. Biochemistry 31(48), 12155–12161 (1992). 30 Appleby TC, Kinsland C, Begley TP, Ealick

SE. The crystal structure and mechanism of orotidine 5´-monophosphate decarboxylase . Proc. Natl Acad. Sci. USA 97(5), 2005–2010, (2000). 31 Toth K, Amyes, TL, Wood BM, Chan K,

Gerlt JA, Richard JP. Product deuterium isotope effects for orotidine 5´-monophosphate decarboxylase: effect of changing substrate and enzyme structure on the partitioning of the vinyl carbanion reaction intermediate. J. Am. Chem. Soc. 132, 7018–7024 (2010). 32 Toth K, Amyes TL, Wood BM, Chan K, Gerlt

JA, Richard JP. Product deuterium isotope effect for orotidine 5´-monophosphate decarboxylase: evidence for the existence of a short-lived carbanion intermediate. J. Am. Chem. Soc. 129, 12946–12947 (2007). 33 Vardi-Kilshtain A, Doron D, Major DT.

Quantum and classical simulations of orotidine monophosphate decarboxylase: support for a direct decarboxylation mechanism. Biochemistry 52(25), 4382–4390 (2013). 34 Amyes TL, Wood BM, Chan K, Gerlt JA,

Richard JP. Formation and stability of a vinyl carbanion at the active site of orotidine 5´-monophosphate decarboxylase: pKa of the C-6 proton of enzyme-bound UMP. J. Am. Chem. Soc. 130, 1574–1575 (2008). 35 Tsang WY, Wood BM, Wong FM et al.

Proton transfer from C-6 of uridine 5´-monophosphate catalyzed by orotidine 5´-monophosphate decarboxylase: formation and stability of a vinyl carbanion intermediate and the effect of a 5-fluoro substituent. J. Am. Chem. Soc. 134, 14580–14594 (2012). 36 Goryanova B, Amyes TL, Gerlt JA, Richard

JP. OMP decarboxylase: phosphodianion binding energy is used to stabilize a vinyl carbanion intermediate. J. Am. Chem. Soc. 133(17), 6545–6548 (2011). 37 Chan KK, Wood BM, Fedorov AA et al.

Mechanism of the orotidine 5´-monophosphate decarboxylase-catalyzed reaction: evidence for substrate destabilization. Biochemistry 48 (24), 5518–5531 (2009). 38 Langley DB, Shojaei M, Chan C et al.

Structure and inhibition of orotidine 5´-monophosphate decarboxylase from Plasmodium falciparum. Biochemistry 47(12), 3842–3854 (2008).

175

Review | Mundra & Kotra 39 Wu N, Pai EF. Crystallographic studies of

51 Sweeney MJ, Davis FA, Gutowski GE, Hamill

native and mutant orotidine 5´ phosphate decarboxylases. Top. Curr. Chem. 238, 23–42 (2004).

RL, Hoffman DH, Poore GA. Experimental antitumor activity of pyrazomycin. Cancer Res. 33(11), 2619–2623 (1973).

40 Amyes TL, Richard JP. Specificity in

transition state binding: the Pauling Model revisited. Biochemistry 52(12), 2021–2035 (2013). 41 Goryanova B, Spong K, Amyes TL, Richard

JP. Catalysis by orotidine 5´-monophosphate decarboxylase: effect of 5-fluoro and 4´-substituents on the decarboxylation of two-part substrates. Biochemistry 52(3), 537–546 (2013). 42 Poduch E, Wei L, Pai EF, Kotra LP.

Structural diversity and plasticity associated with nucleotides targeting orotidine monophosphate decarboxylase. J. Med. Chem. 51(3), 432–438 (2008). nn

Identified the promiscuity existing in the variety of ligands binding to ODCase.

43 Purohit MK, Poduch E, Wei L et al. Novel

cytidine-based orotidine-5´-monophosphate decarboxylase inhibitors with an unusual twist. J. Med. Chem. 55(22), 9988–9997 (2012). 44 Galmarini CM, Popowickz F, Joseph B.

Cytotoxic nucleoside analogs: different strategies to improve their clinical efficacy. Curr. Med. Chem. 15(11), 1072–1082 (2008). 45 Uga H, Kuramori C, Ohta A et al. A new

mechanism of methotrexate action revealed by target screening with affinity beads. Mol. Pharmacol. 70(5), 1832–1839 (2006). 46 Bello AM, Konforte D, Poduch E et al.

Structure–activity relationships of orotidine5´-monophosphate decarboxylase inhibitors as anticancer agents. J. Med. Chem. 52(6), 1648–1658 (2009). 47 Chen JJ, Jones ME. Effect of 6-azauridine on

de novo pyrimidine biosynthesis in cultured Ehrlich ascites cells. Orotate inhibition of dihydrorotase and dihydroorotase dehydrogenase. J. Biol. Chem. 254(11), 4908–4914 (1979). n

One of the very early reports of ODCase inhibitors activity.

48 Cadman EC, Dix DE, Handschumacher RE.

Clinical, biological and biochemical effect of Pyrazofurin. Cancer Res. 38(3), 682–698 (1978). 49 Williams RH, Gerzon K, Hoehn M, Gorman

M, deLong DC. Abstract# MICR 38, 158th American Chemical Society National Meeting. NY, USA, 7–12 September 1969. 50 Christopherson RI, Lyons SD, Wilson PK.

Inhibitors of de novo nucleotide biosynthesis as drugs. Acc. Chem. Res. 35, 961 (2002).

176

52 Gutowski GE, Sweeney MJ, deLong DC,

Hamill RL, Gerzon K. Biochemistry and biological effects of the Pyrazofurins (Pyrazomycins): initial clinical trial. Ann. NY Acad. Sci. 255, 544–551 (1975). 53 Ohnuma T, Roboz J, Shapiro ML, Holland JF.

Pharmacological and biochemical effects of pyrazofurin in humans. Cancer Res. 37(7, Pt 1), 2043–2049 (1977). 54 Cihak A, Veseley J, Skoda J. Azapyrimidine

nucleosides: metabolism and inhibitory mechanism. Adv. Enzyme Regul. 24, 335–354 (1985). 55 Miller BG, Hassell AM, Wolfenden R, Milburn

MV, Short SA. Anatomy of a proficient enzyme: the structure of orotidine 5´monophosphate decarboxylase in the presence and absence of a potential transition-state analog. Proc. Natl Acad. Sci. USA 97(5), 2011–2016 (2000). 56 Poduch E, Bello AM, Tang S, Fujihashi M, Pai

EF, Kotra LP. Design of inhibitors of orotidine monophosphate decarboxylase using bioisosteric replacement and determination of inhibition kinetics. J. Med. Chem. 49(16), 4937–4945 (2006). 57 Smiley JA, Saleh L. Active site probes for yeast

OMP decarboxylase: inhibition constants of UMP and thio-substituted UMP analogues and greatly reduced activity toward CMP-6carboxylate. Bioorg. Chem. 27(4), 297–300 (1999). 58 Gabrielsen B, Kirsi JJ, Kwong CD et al. In vitro

and in vivo antiviral (RNA) evaluation of orotidine 5´-monophosphate decarboxylase inhibitors and analogs including 6-azauridine5´-(ethyl methoxyalaninyl)phosphate (a 5´-monophosphate prodrug). Antiviral Chem. Chemother. 5(4), 209–220 (1994). 59 Smee DF, McKernan PA, Nord LD et al. Novel

pyrazolo[3,4-d]pyrimidine nucleoside analog with broad-spectrum antiviral activity. Antimicrob. Agents. Chemother. 31(10), 1535–1541 (1987). 60 Morrey JD, Smee DF, Sidwell RW, Tseng C.

Identification of active antiviral compounds against a New York isolate of West Nile virus. Antiviral Res. 55(1), 107–116 (2002). 61 Andrei G, de Clercq E. Molecular approaches

for the treatment of hemorrhagic fever virus infections. Antiviral Res. 22(1), 45–75 (1993). 62 Wyde PR, Gilbert BE, Ambrose MW.

Comparison of the anti-respiratory syncytial virus activity and toxicity of papaverine hydrochloride and pyrazofurin in vitro and in vivo. Antiviral Res. 11(1), 15–26 (1989).

Future Med. Chem. (2014) 6(2)

63 de Clercq E. Vaccinia virus inhibitors as a

paradigm for the chemotherapy of poxvirus infections. Clin. Microbiol. Rev. 14(2), 382–397 (2001). 64 Seymour KK, Lyons SD, Phillips L,

Rieckmann, KH, Christopherson RI. Cytotoxic effects of inhibitors of de novo pyrimidine biosynthesis upon Plasmodium falciparum. Biochemistry 33(17), 5268–5274 (1994). 65 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. 18(1), 3–15 (1986). 66 Elgemeie GH, Zaghary WA, Amin KM,

Nasr TM. New trends in synthesis of pyrazole nucleosides as new antimetabolites. Nucleosides Nucleotides Nucleic Acids. 24(8), 1227–1247 (2005). 67 Dorsch HM, Osundwa V, Lam P. Activation

of human B lymphocytes by 8´ substituted guanosine derivatives. Immunol. Lett. 17(2), 125–131 (1988). 68 Wicker LS, Boltz RC Jr, Nichols EA, Miller

BJ, Sigal NH, Peterson LB. Large activated B cells are the primary B-cell target of 8-bromoguanosine and 8-mercaptoguanosine. Cell. Immunol. 106(2), 318–329 (1987). 69 Koo GC, Jewell ME, Manyak CL, Sigal

NH, Wicker LS. Activation of murine natural killer cells and macrophages by 8-bromoguanosine. J. Immunol. 140(9), 3249–3252 (1988). 70 Wicker LS, Ashton WT, Boltz RC Jr et al.

5-halo-6-phenylpyrimidinones and 8-substituted guanosines: biological response modifiers with similar effects on B cells. Cell. Immunol. 112(1), 156–165 (1988). 71 Richard KA, Mortensen RF, Tracey DE.

Cytokines involved in the augmentation of murine natural killer cell activity by pyrimidinones in vivo. J. Biol. Response Mod. 6(6), 647–663 (1987). 72 Skulnick HI, Weed SD, Eidson EE, Renis

HE, Wierenga W, Stringfellow DA. Pyrimidinones. 1. 2-amino-5-halo-6-aryl4(3H)-pyrimidinones. Interferon-inducing antiviral agents. J. Med. Chem. 28(12), 1864–1869 (1985). 73 Brideau RJ, Wolcott JA. Effect of

pyrimidinone treatment on lethal and immunosuppressive murine cytomegalovirus infection. Antimicrob. Agents Chemother. 28(4), 485–488 (1985). 74 Wierenga W. Antiviral and other bioactivities

of pyrimidinones. Pharmacol. Ther. 30(1), 67–89 (1985).

future science group

Design of inhibitors of ODCase 75 Goodman MG, Hennen WJ. Distinct effects

of dual substitution on inductive and differentiative activities of C8-substituted guanine ribonucleosides. Cell. Immunol. 102, 395–402 (1986).

pyrimidine biosynthesis upon Plasmodium falciparum. Biochemistry 33(17), 5268–5274 (1994). 83 Yadav MK, Pandey SK, Swati D. Drug target

prioritization in Plasmodium falciparum through metabolic network ana­lysis, and inhibitor designing using virtual screening and docking approach. J. Bioinform. Comput. Biol. 11(4), 1350003 (2013).

76 Reller ME, Chen WH, Dalton J, Lichay MA,

Dumler JS. Multiplex 5´ nuclease quantitative real-time PCR for clinical diagnosis of malaria and species-level identification and epidemiologic evaluation of malaria-causing parasites, including Plasmodium knowlesi. J. Clin. Microbiol. 51(9), 2931–2938 (2013).

80 Scheibel LW, Scherman IW. Malaria:

Identification of inhibitors for putative malaria drug targets among novel antimalarial compounds. Mol. Biochem. Parasitol. 175(1), 21–29 (2011). 82 Seymour KK, Lyons SD, Phillips L,

Rieckmann KH, Christopherson RI. Cytotoxic effects of inhibitors of de novo

future science group

Antimalarial activities of 6-iodouridine, its prodrugs and potential for combination therapy. J. Med. Chem. 56(6), 2348–2358 (2013). S, Fairlamb AH. Trypanosoma brucei (UMP synthase null mutants) are avirulent in mice, but recover virulence upon prolonged culture in vitro while retaining pyrimidine auxotrophy. Mol. Microbiol. 90(2), 443–455 (2013).

nn

Identified the first covalent inhibitor to ODCase, introducing new paradigm into the binding site residues and their function.

85 Bello AM, Poduch E, Liu Y et al. Structure-

activity relationships of C6-uridine derivatives targeting Plasmodia orotidine monophosphate decarboxylase. J. Med. Chem. 51(3), 439–448 (2008). 86 Fujihashi M, Wei L, Kotra LP, Pai EF.

Structural characterization of the molecular events during a slow substrate-product transition in orotidine 5´-monophosphate decarboxylase. J. Mol. Biol. 387(5), 1199–1210 (2009).

Principles and Practice of Malariology. Wernsdorfer WH, McGreggor I (Eds). Churchill Livingstone, London, UK, 234–242 (1988). 81 Crowther GJ, Napuli AJ, Gilligan JH et al.

89 Crandall IE, Wasilewski E, Bello AM et al.

90 Ong HB, Sienkiewicz N, Wyllie S, Patterson

79 Sherman IW. Biochemistry of Plasmodium

(malarial parasites). Microbiol. Rev. 43(4), 453–495 (1979).

in Drug Discovery: A Guide for Medicinal Chemists and Pharmacologists. John Wiley and Sons. Inc., Hoboken, NJ, USA (2005).

potent, covalent inhibitor of orotidine 5´-monophosphate decarboxylase with antimalarial activity. J. Med. Chem. 50(5), 915–921 (2007).

78 van Dyke K, Tremblay GC, Lantz CH,

Szustkiewicz, C. The source of purines and pyrimidines in Plasmodium berghei. Am. J. Trop. Med. Hyg. 19(2), 202–208 (1970).

88 Copeland RA. Evaluation of Enzyme Inhibitors

84 Bello AM, Poduch E, Fujihashi M et al. A

77 Goh XT, Lim YA, Vythilingam I et al.

Increased detection of Plasmodium knowlesi in Sandakan division, Sabah as revealed by PlasmoNex™. Malar. J. 12, 264 (2013).

| Review

87 Fujihashi M, Bello AM, Poduch E et al. An

unprecedented twist to ODCase catalytic activity. J. Am. Chem. Soc. 127, 15048–15050 (2005). nn

91 Lewis CA, Wolfenden R. Indiscriminate

binding by orotidine 5´-phosphate decarboxylase of uridine 5´-phosphate derivatives with bulky anionic C6 substituents. Biochemistry 46(46), 13331–13343 (2007). „„Patents 101 Williams RH, Hoehn MM: US3674774

(1972). 102 Kotra LP, Pai EF, Bello AM, Fujihashi M:

US8067391 (2009). „„Website 201 Protein database.

www.rcsb.org/pdb/home/home.do

Catalytic promiscuity of ODCase promoting biochemical reactions other than decarboxylation.

www.future-science.com

177