+ All Categories
Home > Documents > AAC Accepts, published online ahead of print on 29 March...

AAC Accepts, published online ahead of print on 29 March...

Date post: 25-Apr-2018
Category:
Upload: dangquynh
View: 218 times
Download: 4 times
Share this document with a friend
36
Preclinical evaluation of antifolate QN254 (5-chloro-N*6*-(2,5-dimethoxy-benzyl)- 1 quinazoline-2,4,6-triamine) as antimalarial drug candidate 2 3 Alexis Nzila 1,2,* , Matthias Rottmann 3 , Penchit Chitnumsub 4 , Stevens M. Kiara 1 , Sumalee 4 Kamchonwongpaisan 4 , Cherdsak Maneeruttanarungroj 4 , Supannee Taweechai 4 , Bryan K.S. 5 Yeung 5 , Anne Goh 5 , Suresh B. Lakshminarayana 5 , Bin Zou 5 , Josephine Wong 5 , Ngai Ling 6 Ma 5 , Weaver Margaret 6 , Thomas H. Keller 5 , Veronique Dartois 5 , Sergio Wittlin 3 , Reto Brun 3 , 7 Yongyuth Yuthavong 4 and Thierry T. Diagana 5 8 9 10 1 Kenya Medical Research Institute (KEMRI)/Wellcome Trust Collaborative Research 11 Program, PO Box 230, 80108, Kilifi, Kenya; 2 University of Oxford, Nuffield Department of 12 Medicine, John Radcliffe Hospital, Oxford, UK; 3 Swiss Tropical Institute (STI) Parasite 13 Chemotherapy, Socinstrasse 57, PO Box, CH-4002 Basel, Switzerland; 4 National Centre for 14 Genetic Engineering and Biotechnology (BIOTEC), 113 Thailand Science Park, NSTDA, 15 Pathumthani 12120, Thailand; 5 Novartis Institute for Tropical Diseases (NITD), 10 Biopolis 16 Road #05-01 Chromos, 138670 Singapore; 6 Novartis Institute for Biomedical Research 17 (NIBR), 250 Massachusetts Ave, Cambridge, MA 02139, USA 18 19 * Corresponding author: Dr Alexis Nzila, Kenya Medical Research Institute 20 (KEMRI)/Wellcome Trust Collaborative Research Program, PO Box 230, 80108 Kilifi, Kenya. 21 Tel: +254-417-522-535; fax: +254-417-522-290; e-mail: [email protected] 22 23 Copyright © 2010, American Society for Microbiology and/or the Listed Authors/Institutions. All Rights Reserved. Antimicrob. Agents Chemother. doi:10.1128/AAC.01526-09 AAC Accepts, published online ahead of print on 29 March 2010 on June 9, 2018 by guest http://aac.asm.org/ Downloaded from
Transcript

Preclinical evaluation of antifolate QN254 (5-chloro-N*6*-(2,5-dimethoxy-benzyl)-1

quinazoline-2,4,6-triamine) as antimalarial drug candidate 2

3

Alexis Nzila1,2,*, Matthias Rottmann3, Penchit Chitnumsub4, Stevens M. Kiara1, Sumalee 4

Kamchonwongpaisan4, Cherdsak Maneeruttanarungroj4, Supannee Taweechai4, Bryan K.S. 5

Yeung5, Anne Goh5, Suresh B. Lakshminarayana5, Bin Zou5, Josephine Wong5, Ngai Ling 6

Ma5, Weaver Margaret6, Thomas H. Keller5, Veronique Dartois5, Sergio Wittlin3, Reto Brun3, 7

Yongyuth Yuthavong4 and Thierry T. Diagana5 8

9

10

1Kenya Medical Research Institute (KEMRI)/Wellcome Trust Collaborative Research 11

Program, PO Box 230, 80108, Kilifi, Kenya; 2University of Oxford, Nuffield Department of 12

Medicine, John Radcliffe Hospital, Oxford, UK; 3Swiss Tropical Institute (STI) Parasite 13

Chemotherapy, Socinstrasse 57, PO Box, CH-4002 Basel, Switzerland; 4National Centre for 14

Genetic Engineering and Biotechnology (BIOTEC), 113 Thailand Science Park, NSTDA, 15

Pathumthani 12120, Thailand; 5Novartis Institute for Tropical Diseases (NITD), 10 Biopolis 16

Road #05-01 Chromos, 138670 Singapore; 6Novartis Institute for Biomedical Research 17

(NIBR), 250 Massachusetts Ave, Cambridge, MA 02139, USA 18

19

*Corresponding author: Dr Alexis Nzila, Kenya Medical Research Institute 20

(KEMRI)/Wellcome Trust Collaborative Research Program, PO Box 230, 80108 Kilifi, Kenya. 21

Tel: +254-417-522-535; fax: +254-417-522-290; e-mail: [email protected] 22

23

Copyright © 2010, American Society for Microbiology and/or the Listed Authors/Institutions. All Rights Reserved.Antimicrob. Agents Chemother. doi:10.1128/AAC.01526-09 AAC Accepts, published online ahead of print on 29 March 2010

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

2

1 Abstract 24

Drug resistance against dihydrofolate reductase (DHFR) inhibitors – such as pyrimethamine 25

(PM) – has now spread to almost all malaria endemic regions, rendering antifolate-based 26

malaria treatments highly ineffective. We have previously shown that the di-amino quinazoline 27

QN254 is active against the highly pyrimethamine-resistant Plasmodium falciparum V1S 28

strain, suggesting that QN254 could be used to treat malaria in regions with a high prevalence 29

of antifolate resistance. Here, we further demonstrate that QN254 is highly active against 30

Plasmodium falciparum clinical isolates, displaying various levels of antifolate drug resistance, 31

and we provide biochemical and structural evidence that QN254 binds and inhibits the function 32

of both the wild-type and quadruple-mutant form (V1S) of the DHFR enzyme. In addition, we 33

have assessed QN254 oral bioavailability, efficacy and safety in vivo. The compound displays 34

favorable pharmacokinetic properties following oral administration in rodents. The drug was 35

remarkably efficacious against Plasmodium berghei and could fully cure infected mice with 36

three daily oral doses of 30 mg/kg. In the course of these efficacy studies, we have uncovered 37

some dose limiting toxicity at higher doses that was confirmed in rats. Thus, despite its relative 38

in vitro selectivity toward the Plasmodium DHFR enzyme, QN254 does not show the adequate 39

therapeutic index to justify its further development as a single agent. 40

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

3

41

2 Introduction 42

Malaria control is a global public health priority that has been hampered by the rapid 43

development and spread of resistance against antimalarials. As a consequence, the WHO 44

recommends the use of artemisinin-containing combination therapies (ACTs) as first-line 45

treatment for malaria. Although ACTs are designed to reduce the chance of artemisinin drug 46

resistance development, there are considerable concerns that this may already have occurred. 47

For instance, there is now mounting evidence that the efficacy of artemisinin derivatives is 48

reduced in South-East Asia, where artemisinin derivatives have been used for a long time as 49

monotherapies (7, 28, 53). This is a cause of concern since the spread of artemisinin resistance 50

will compromise the usefulness of ACTs globally. Thus there is an urgent need to discover and 51

develop new alternative drugs. 52

For several decades, dihydrofolate reductase (DHFR) has been targeted with different classes 53

of chemical entities for the development of new therapies for a broad range of therapeutic 54

indications including several parasitic diseases (13). DHFR catalyzes the reduction of 55

dihydrofolate (DHF) to tetrahydrofolate (THF), which is an essential cofactor in the 56

biosynthesis of deoxythymidylate monophosphate (dTMP), a metabolite essential to DNA 57

synthesis and cell replication. 58

Pyrimethamine (PM) is a potent inhibitor of the Plasmodium DHFR enzyme and this 59

compound has been widely used in combination with the dihydropteroate synthetase (DHPS) 60

inhibitor sulfadoxine. Unfortunately, pyrimethamine resistance is now common, rendering this 61

drug ineffective (30). One of the documented mechanisms of antifolate resistance is through 62

the mutation of the target itself – the pfdhfr-ts gene, encoding for a bi-functional enzyme which 63

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

4

possesses both DHFR and thymidylate synthase (TS) activities carried by two distinct sub-64

domains (29). Pyrimethamine resistance is associated with the mutation of the amino acid Ser 65

to Asn at codon 108 of DHFR (S108N). Ancillary mutations of N51I and C59R are associated 66

with an increase in resistance, and the presence of the mutation I164L results in an even higher 67

level of PM resistance (14, 47). The presence of these mutations significantly decreases the 68

sensitivity of the PfDHFR enzyme to pyrimethamine inhibition in a biochemical assay (48). 69

The antifolate triazine WR99210 (3, 22) is potent against Plasmodium falciparum bearing 70

quadruple mutations of DHFR at S108N, N51I, C59R, and I164L (QM PfDHFR) (20). 71

However, WR99210 has shown limited efficacy in vivo due to poor oral bioavailability and 72

displayed some gastrointestinal toxicity. Attempts have been made to circumvent these issues 73

and a pro-drug form of WR99210 known as PS-15 has been shown to be orally active (2). The 74

resolution of the three-dimensional structure of wild-type (WT) and QM PfDHFR – with either 75

PM or WR99210 bound to its active site – provided structural insights into DHFR 76

pyrimethamine resistance mechanisms as well as some understanding of the structural features 77

of WR99210 that allow this compound to retain affinity for QM PfDHFR (55, 56). 78

The quinazoline (QN) pharmacophore has been successfully employed to design drugs for the 79

treatment of cancer and other human diseases. For example, the DHFR inhibitor, trimetrexate 80

(5-methyl-6-[(3,4,5-trimethoxy-phenylamino)-methyl]-quinazoline-2,4-diamine) has been 81

developed to treat various cancers in human patients (15, 32, 33). In a previous study, a series 82

of quinazoline derivatives was tested against the highly pyrimethamine-resistant Plasmodium 83

falciparum strain (V1S) and the DHFR inhibitor 2,4-diamino-5-chloro-6-[N-(2,5-84

dimethoxybenzyl)-amino]quinazoline (or QN254 in this paper and compound 1 in (34)) was 85

found to have potent activity – with an IC50 value (inhibitory concentration that reduces 86

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

5

parasite growth by 50% in vitro) of 9 nM (34). Considering the increasingly widespread 87

pyrimethamine drug resistance, we set out to perform the experiments described in this paper 88

with the goal of further assessing the potential of QN254 as a candidate antimalarial to replace 89

the failing pyrimethamine. Collectively, our data demonstrate that QN254 (i) binds and inhibits 90

the QM PfDHFR enzyme, (ii) is active on drug-resistant clinical isolates and (iii) displays 91

pharmacological properties compatible with an oral antimalarial drug candidate. However, 92

preliminary toxicological findings indicate that QN254 does not show a therapeutic window 93

sufficiently large to warrant its progression to the next development stage. 94

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

6

3 Material and Methods 95

Drugs and reagents 96

Pyrimethamine (PM) and chloroquine (CQ) used as standard drugs for evaluation on clinical 97

isolates were purchased from Sigma (Cole, UK). WR99210 was a gift from Professor Steve 98

Ward (Liverpool Tropical School, UK). In the Ki determination, cycloguanil (CGL) and 99

WR99210 were gifts from Professor Tirayut Vilaivan (Chulalongkorn University, Thailand). 100

Standard drugs for in vivo efficacy testing were obtained from Mepha Ltd, Switzerland 101

(artesunate [ART]), Sigma, USA (chloroquine diphosphate) and F. Hoffmann-LaRoche Ltd, 102

Switzerland (mefloquine hydrochloride). QN254 was prepared as described elsewhere (39). 103

DHFR biochemical assay and X-ray structure determination 104

Ki determination: Recombinant PfDHFR enzymes of P. falciparum and human were prepared 105

from E. coli BL21(DE3) bearing pET17b expression plasmids of P. falciparum wild-type (WT) 106

and quadruple mutant (S108N, N51I, C59R, I164L) [QM pfdhfr] (17), and human dhfr (hdhfr) 107

(18), using a methotrexate column as described previously (51). Binding affinities (Ki values) 108

of QN254 and standard antifolate antimalarials, cycloguanil, pyrimethamine and WR99210 109

were determined as described elsewhere (51). 110

X-Ray structure determination 111

Preparation of recombinant PfDHFR-TS bifunctional enzymes was carried out as previously 112

reported (4). Crystals were flash-frozen in liquid nitrogen by dipping for 10 s in a 113

corresponding crystallization solution containing 20% glycerol as a cryoprotectant. X-ray 114

diffraction data were collected under a cold nitrogen stream (100 K) at a wavelength of 1.54 Å 115

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

7

on an FR591 rotating anode X-ray generator equipped with a nonius KappaCCD detector. Data 116

were processed using DENZO and SCALEPACK in the HKL2000 suite (35). Refinement was 117

performed using a 2.1 Å 1J3K crystal structure as a template in both CNS (1) and REFMAC5 118

(26) along with model building in program O (16) and model validation in PROCHECK (24, 119

25) and Moleman2 (21). Figures were prepared with PyMOL (6). 120

Plasmodium strains and culturing methods 121

3.1.1 Parasite strains and isolates 122

We analyzed the in vitro activity of QN254 against clinical isolates from Kenya with different 123

dhfr genotypes. These clinical isolates were collected in Kilifi between 2006 and 2008 as part 124

of the malaria studies, and were in vitro adapted for long-term culture as described elsewhere 125

(40). 126

3.1.2 In vitro culture and chemosensitivity test 127

P. falciparum cultures were carried out in RPMI 1640 (GIBCO BRL, UK) medium 128

supplemented with 10% (v/v) normal human serum, 25 mM bicarbonate, 2 mM glutamine, 25 129

mM HEPES buffer, physiological concentrations of para-aminobenzoic acid (5 nM) and folic 130

acid (23 nM). Antimalarial activity was measured in the presence of varying concentrations of 131

each compound using radioisotopic incorporation (49). Results were expressed as the drug 132

concentration required for 50% inhibition (IC50) of [3H]hypoxanthine incorporation into 133

parasite nucleic acid, using non-linear regression analysis of the dose–response curve. 134

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

8

3.1.3 Parasite genotyping of dhfr 135

After in vitro adaptation of parasites, infected blood samples were spotted onto filter paper and 136

stored. Parasite genomic material from these filter papers was prepared using the methanol 137

procedure and point mutations at codons 108, 51, 59 and 164 of dhfr were analyzed by PCR 138

and restriction enzyme digestion (PCR-RFLP) as described elsewhere (31). Statistical analyses 139

were carried out using Stata, version 9 (StataCorp, College Station, TX, USA), using the 140

Kruskal–Wallis non-parametric test for comparison of medians. The level of significance was 141

set at p<0.05. 142

In vivo pharmacokinetic studies 143

The Institutional Animal Care and Use Committee (IACUC) of Novartis Institute for Tropical 144

Diseases, registered with the Agri-Food and Veterinary Authority (AVA), Government of 145

Singapore, approved all animal experimental protocols. QN254 was formulated at a 146

concentration of 2.5 mg/ml for a dose of 25 mg/kg given orally (p.o.) and at 1 mg/ml 147

concentration for a dose of 5 mg/kg given intravenously (i.v.).The solution formulation for i.v. 148

dosing contained 10% NMP (n-methyl pyrrolidone), 30% PEG 400 and 60% of 10% Vitamin 149

E-TPGS. The suspension formulation for p.o dosing was 0.5% CMC (carboxymethyl 150

cellulose). For PK studies, blood samples were collected at several time points following p.o. 151

and i.v. dosing in mice and rats. 152

3.1.4 Extraction and LCMS analysis of QN254 153

Plasma samples were extracted with acidified acetonitrile for QN254 using an extractant to 154

plasma ratio of 8:1. Analyte quantification was performed by LC/MS/MS. Liquid 155

chromatography was performed using an Agilent 1100 HPLC system (Santa Clara, CA, USA), 156

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

9

with an Agilent Zorbax XDB Phenyl (3.5µ, 4.6 × 75 mm) column at an oven temperature of 157

35°C, coupled with an API3200 triple quadruple mass spectrometer (Applied Biosystems, 158

Foster City, CA, USA). Instrument control and data acquisition were performed using Analyst 159

1.4.2 software (Applied Biosystems). The mobile phases used were A: water–acetic acid 160

(99.8:0.2, v/v) and B: acetonitrile–acetic acid (99.8:0.2, v/v), using a gradient, with flow rate of 161

1.0 ml/min and run time of 5 min. Under these conditions, the analyte retention time was 2.9 162

min. Compound detection on the mass spectrometer was performed in electrospray positive 163

ionization mode and using multiple reaction monitoring (MRM) for specificity (transitions 164

360.3/209.2 and 360.3/174.2) together with their optimized MS parameters. Analysis of study 165

samples was performed on different days using an identical method with a ten-point calibration 166

curve spanning three orders of magnitude in compound concentration. The lower and upper 167

limits of quantification were 12 ng/ml and 7500ng/mL respectively. Intra-day variability was 168

established with triplicate quality control samples at three concentration levels. Relative 169

standard deviation was under 15%. Inter-day variability is not reported as calibration curves 170

were re-prepared and reanalyzed with every set of study samples and results accepted if inter-171

day variability was found to be under 15%. 172

3.1.5 Pharmacokinetic data analysis 173

The mean value from three animals at each time point was plotted against time to give the 174

plasma concentration–time profile. Pharmacokinetic parameters were determined using 175

WinNonlin Professional, version 5.0.1 (Pharsight, CA, USA), by non-compartmental modeling 176

using software model 200 for oral dosing and model 201 for intravenous dosing. The oral 177

bioavailability (F) was calculated as the ratio between the area under the curve (AUC) 178

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

10

following oral administration and the AUC following intravenous administration corrected for 179

dose (F = AUC p.o. × dose i.v./AUC i.v. × dose p.o.). 180

In vivo antimalarial efficacy studies 181

All in vivo efficacy studies were carried out at the Swiss Tropical Institute, adhering to the 182

regulations of the veterinary authorities of Basel-Stadt. The murine P. berghei ANKA was 183

used as previously described (37). Experimental and control groups of 5 female NMRI mice 184

(20–22 g) were infected intravenously with 1 × 108 parasitized erythrocytes/ml, in a volume of 185

0.2 ml (from a fresh mouse donor). In untreated control mice, parasitemia rises regularly to 186

~30% by day 3 post-infection and causes death of the animals between day 5 and day 7 post-187

infection. 188

Experimental compounds were prepared in 0.5% carboxymethyl cellulose (CMC) and 189

administered orally. Comparators (Artesunate, Chloroquine and Mefloquine) were prepared in 190

ETPGS vehicle (10% EtOH, 30% PEG400 and 60% of a 10% Vitamin ETPGS solution) and 191

administered orally. Parasitemia was determined microscopically. Activity was calculated as 192

the difference between parasitemia for the control and treated groups expressed as a per cent 193

relative to the control group. The survival time was also recorded up to 30 days after infection. 194

A compound was considered curative if the animal survived to day 30 after infection with no 195

detectable parasites. The ED50 and ED90 values were assessed at day 3 by plotting the dose as a 196

function of parasitemia, using non-linear fitting with the Microcal Origin Statistical Program 197

(OriginLab, Northampton, MA, USA). All data reported are based on n≥10 mice, except for 198

the three times treatment with QN254 which, because of the observed toxicity, were performed 199

only once with a cohort of 5 mice. 200

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

11

Rat toxicological study 201

QN254 was suspended in 10 % of a 1% (v/v) aqueous Solutol HS15 solution and 90% of a 202

0.5 % aqueous Methylcellulose M0555 solution and administered to groups of 5 male rats 203

at daily oral (by gavage) doses of 50, 150 or 500 mg/kg/day. Animals (Wistar rats, Harlan 204

Laboratories Ltd., Füllinsdorf, Switzerland) were approximately 10 weeks of age (254 to 205

292 g) at the start of dosing. Clinical observations, body weight and food consumption 206

determinations, clinical pathology (hematology and clinical chemistry) evaluations, gross 207

pathology examinations without organ weight determinations (due to premature sacrifice 208

of all treated groups) were performed on all groups. Microscopic examinations were 209

conducted on all gross lesions and on a limited standard list of organs and tissues from 210

animals assigned to the control and low-dose groups. At day 1 and day 8 blood samples 211

were collected for toxicokinetic analyses. 212

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

12

4 Results 213

Inhibitory activity (Ki): QN254 is a potent inhibitor of both the wild-type and quadruple 214

mutant plasmodium DHFR enzymes 215

QN254 was compared to the antimalarials pyrimethamine (PM) and cycloguanil (CGL) – 216

inactive on QM PfDHFR – as well as the triazine WR99210, a potent inhibitor of all mutant 217

PfDHFR, including QM PfDHFR (20). The results of these experiments are summarized in 218

Table 1. Consistent with previously published results showing that QN254 is active against the 219

highly pyrimethamine-resistant strain V1S, our results show that QN254 displays binding 220

affinities comparable to WR99210 against both WT PfDHFR (Ki= 0.39 ± 0.05 nM) and QM 221

PfDHFR (Ki=0.58 ± 0.06 nM). It is also worth noting that QN254 appears to show a slightly 222

better selectivity than WR99210 as shown by the higher human/plasmodium Ki ratio (26 for 223

QN254 vs 15 for WR99210). 224

WR99210 and QN254 have similar binding modes on QM PfDHFR 225

To determine the structural features important for QN254 binding to PfDHFR, we solved the 226

crystal structure of a QN254/QM PfDHFR-TS (V1S) complex at 2.7 Å resolution 227

(Supplementary Table S1). Figure 1 describes the structure of this complex and shows the key 228

interactions required for QN254 binding. QN254 makes direct interactions with several key 229

residues in the DHFR active site as well as interactions via a water-mediated hydrogen bond. 230

Compared to WR99210 binding (56), the diamino-quinazoline scaffold of QN254 hydrogen 231

bonds with the side chain of Asp54 and gains an additional H-bond with Thr185 to replace the 232

loss of an H-bond with Tyr170 presumably due to the larger nature of the quinazoline ring 233

(Figure 1A and 1B). Extensive hydrogen bonds between the enzyme carbonyl backbones at 234

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

13

Ile14, Cys15, Leu164 and the diamino-quinazoline ring were established as observed in the 235

WR99210 structure. Interactions which increase the binding of QN254 over WR99210 include 236

a water-mediated hydrogen bond between the N-6 of the quinazoline ring with both the side 237

chain of Asn108 and the C-2 hydroxyl of the nicotinamide adenosine dinucleotide phosphate 238

(NADPH) ribose (Figure 1A). In the WR99210 structure, a water molecule also mediates a 239

hydrogen bond between Ser111 and the ribose of NADPH (Figure 1B), but in this case, only 240

enhances the interaction with NADPH (56). QN254 makes additional van der Waals contacts 241

and π−π interactions with Leu46 and Met55 and Phe58 side chains. The phenyl substituent of 242

both inhibitors makes similar van der Waals interactions with Ile112, Pro113, Phe116 and 243

Leu119. Consistent with the biochemical data reported above, comparison of the QN254/QM 244

PfDHFR and WR99210/QM PfDHFR complexes (Figure 1C) shows that both compounds 245

overlap extensively within the active site of the QM PfDHFR enzyme. It appears that, in both 246

cases, the flexibility of the linker region between the diamino-quinazoline or diamino-247

dihydrotriazine moieties as well as the lipophilicity of the phenyl ring are critically important 248

for binding to the QM PfDHFR. Moreover, QN254 takes advantage of the mutation at Asn108 249

by making an additional interaction via a water molecule (Figure 1A: 6-N…O, 3.22 Å) – 250

absent in the WR99210/V1S structure – which presumably contributes to its better affinity for 251

the QM enzyme compared with WR99210. 252

QN254 shows potent antimalarial activity against P. falciparum pyrimethamine drug-253

resistant clinical isolates with various PfDHFR genotypes 254

To further demonstrate that QN254 is active against P. falciparum even in areas of widespread 255

PM drug resistance, we determined the in vitro activity of QN254 against 27 clinical isolates 256

collected in Kilifi (Kenya) between 2006 and 2008. These isolates were part of a larger study 257

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

14

in which we measured the activity of standard antifolate drugs (including PM) (19). For 258

comparison purposes, in Table 2, we report again the data pertaining to PM along with the 259

previously published results obtained with QN254 (19). As expected, PM was found to be 260

largely inactive on these clinical isolates, with a median IC50 of 733.26 nM. In agreement with 261

these data, and as discussed in our previous report, no wild-type isolates were found and more 262

than 72% of the tested isolates carried triple mutations in pfdhfr (S108N+N51I+C59R). 263

Interestingly, one isolate carried the four mutations (S108N+N51I+C59R+I164L), mutations 264

present in the V1S form (19). 265

In sharp contrast to the results obtained with PM, QN254 was potent against all isolates, with a 266

median IC50 value of 9.55 nM. The median IC50 values of QN254 against double mutants 267

(S108N+N51I or S108N+C59R) and triple mutants were 4.48 nM and 11.66 nM, respectively, 268

and this difference was not statistically significant (Table 2). The quadruple mutant isolate had 269

an IC50 of 7.61 nM. Collectively, these data clearly show that QN254 is active on clinically 270

relevant P. falciparum isolates with a wide range of mutations in pfdhfr (including quadruple 271

mutations) associated with significant PM drug resistance. 272

QN254 is orally bioavailable and displays slow but complete absorption as well as a long 273

half-life in rodents 274

To determine the pharmacokinetic (PK) properties of QN254 in vivo, we measured the plasma 275

concentration–time profile upon oral and intravenous administration in both mice and rats 276

(Figure 2). Pharmacokinetic parameters are reported in Supplementary Table S2. In both rodent 277

species, QN254 displayed a high volume of distribution, greater than total body water (volume 278

of distribution at steady state (Vss)=2.85 and 4.7 l/kg in mice and rats, respectively), and the 279

total systemic clearance was moderate to low at 57% and 30% of hepatic blood flow in mice 280

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

15

and rats, respectively. Consistently, QN254 displayed an intravenous (i.v.) half-life of 0.8 h in 281

mice and a relatively longer half-life of 3.5 h rats. Following oral administration, the 282

elimination half-life was around 6 h and 9 h in mice and rats, respectively. QN254 oral 283

absorption was slow but complete with a time to maximum concentration (Tmax) of 4 h and 284

apparent oral bioavailability of 100% in mice. These data indicate that the low aqueous 285

solubility of the compound (~2 mg/l at pH 6.8) does not limit its absorption in rodents. In rats, 286

QN254 absorption is similarly slow and the relative oral bioavailability is 63%. Notably 287

maximum plasma concentrations (Cmax) values were about 5.1 and 2.8 µM in mice and rats, 288

respectively, which are concentrations several hundred fold above the P. falciparum IC50 value. 289

QN254 is orally efficacious in the P. berghei infected malaria mouse model 290

We tested the efficacy of QN254 in the P. berghei malaria mouse model, a reference animal 291

model (12). Upon oral administration of single doses at 10, 15, 30, 60 and 100 mg/kg, QN254 292

showed a dose-dependent parasitemia reduction, 59% and 73% for 10 and 15 mg/kg doses, 293

respectively; >99.99% for all other doses (Table 3). The efficacious doses of QN254 required 294

to inhibit the growth of 50% (ED50) and 90% parasitemia (ED90) were 7.2 and 12.4 mg/kg, 295

respectively, values that are comparable to those of artesunate (5.9 and 20.5 mg/kg) but slightly 296

higher than mefloquine (3.8 and 5.2 mg/kg) and chloroquine (1.9 and 4.2 mg/kg) (52). 297

These results are consistent with the good pharmacokinetic (PK) data by the oral route reported 298

above, however one should bear in mind that the PK analysis was done in naive mice and one 299

cannot rule out that the PK parameters in infected mice might be different. Mouse survival was 300

significantly increased to 8.1 days with a single dose treatment at 30 mg/kg and to 12 days at 301

60 mg/kg. At 100 mg/kg, QN254 displayed activity superior to the currently marketed 302

antimalarial drugs artesunate and chloroquine, with parasitemia reduction >99.99% and an 303

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

16

average mouse survival prolongation of 23.4 days. In addition, 50% of mice were cured with a 304

single dose of 100 mg/kg QN254, while no mice were cured with a similar drug treatment 305

regimen for all three standard antimalarial drugs. 306

At three daily oral doses of 30 mg/kg, the cure rate was improved to 80% with a mouse 307

survival prolongation of about 28.4 days. We found some dose-limiting toxicity at 3 × 60 and 3 308

× 100 mg/kg as six out of the 10 treated mice died at around day 10, despite being free of 309

parasites at the time. The four surviving mice were cured and free of parasites at day 30. 310

Collectively, our data show that QN254 displays potent in vivo oral activity in the P. berghei 311

mouse model and cured infected mice upon administration of three daily oral doses as low as 312

30 mg/kg. However, three-day dosing at higher doses (≥ 60 mg/kg) led to toxicity and death in 313

some animals. These results indicate that the therapeutic window of QN254 might be too 314

narrow. 315

QN254 displays severe toxicity in rats and has a small therapeutic window 316

In an exploratory 2-week rat toxicology study we found that QN254 was not tolerated upon 317

repeated oral administration of daily dose ≥ 50 mg/kg. In this study, all animals had to be 318

euthanized by day eight, because of serious adverse effects, body weight loss, decreased food 319

consumption and other clinical signs (soft feces, piloerection, reduced motor activity). 320

In all rats sacrificed preterm, histopathological analysis revealed marked gastrointestinal tract 321

and bone marrow toxicity. The major findings included degenerative/regenerative, atrophic and 322

inflammatory changes in the gastric tract—mostly in the small intestine and cecal mucosa—as 323

well as massive bone marrow atrophy. The affected tissues being very proliferative, this type of 324

toxicity is consistent with on-target effects through sustained inhibition of DHFR and 325

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

17

inhibition of cell proliferation as it has been previously reported for other DHFR inhibitors (2, 326

54). 327

Toxicokinetics data generated in the course of this study showed that in rats, upon 328

administration of a 50 mg/kg dose of QN254, daily plasma exposure at day 1 (AUC0-24h= 329

27,600 ng*h/mL) is only about 2 times the plasma exposure level reached at the efficacious 330

dose in mice (AUC0-24h≈ 16,990 ng*h/mL at the ED99=26.5 mg/kg). Thus QN254 therapeutic 331

window is smaller than 2. 332

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

18

333

5 Discussion 334

We have previously shown that QN254 is active in vitro against the P. falciparum strain (V1S) 335

carrying a copy of the QM-pfdhfr gene. Here, we provide data that the diamino-quinazoline 336

DHFR inhibitor QN254 (i) binds to and inhibits the QM PfDHFR enzyme, (ii) displays potent 337

antimalarial activity against antifolate drug-resistant P. falciparum clinical isolates with 338

various DHFR genotypes, including QM-pfdhfr, (iii) shows pharmacokinetic properties 339

compatible with oral dosing and (iv) displays excellent in vivo therapeutic efficacy in the P. 340

berghei malaria mouse model. 341

Because antifolate drugs such as pyrimethamine and proguanil have lost their efficacy due to 342

resistance, it is crucial for any new molecule targeting this enzyme to show activity against all 343

pyrimethamine-resistant DHFR mutant forms. In an earlier study, we assessed the in vitro 344

activity of WR99210 against the same clinical isolates tested in this publication (mean IC50 of 345

about 0.72 nM (19)). Thus when compared to QN254, WR99210 is about 10 times more potent 346

against parasite despite having a slightly lower binding affinity to PfDHFR than QN254. It 347

remains to be determined whether better diffusion across membranes of parasitized red blood 348

cells and/or an additional off-target effect could explain WR99210 higher cellular potency. 349

Using a functional/biochemical assay and crystallographic structural data, we provide further 350

evidence that QN254 binds to and inhibits Plasmodium WT and QM DHFR enzymes 351

efficiently. The resolution of the three-dimensional structure of wild-type (WT) and QM 352

PfDHFR – with either PM or WR99210 bound to its active site – provided structural insights 353

into DHFR pyrimethamine resistance mechanisms as well as some understanding of the 354

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

19

structural features allowing WR99210 to retain affinity for QM PfDHFR (55, 56). Our 355

crystallographic data, describing the structure of a ternary complex of QN254 bound to QM 356

PfDHFR, show that QN254 binds to its target in a manner similar to WR99210. Previous X-ray 357

crystallographic studies (5) with a compound closely related to QN254 and the Pneumocystis 358

carinii DHFR enzyme suggested that Phe 69 (P. carinii numbering or Phe 116 in Pf 359

numbering) was engaged in close hydrophobic contact with the 5′ methoxy group on the 360

phenyl ring of the lipophilic tail. The authors also speculated that this interaction may 361

contribute to the relative selectivity toward P. carinii (and possibly P. falciparum) as the 362

conserved phenylalanine was not present in the mammalian DHFR enzymes. Careful analysis 363

of our co-crystal QN254/PfDHFR structure failed to reveal such an interaction and in fact we 364

could only identify weak van der Waals interactions between the lipophilic tail and the 365

hydrophobic cavity present in the active site of QM PfDHFR. The structure activity 366

relationship (SAR) data for the class of quinazoline compounds are not very clear and are 367

currently limited to what is described in previous publications (8-11, 38, 39). 368

Consistent with our structural data demonstrating the absence of specific interactions between 369

the phenyl ring tail and PfDHFR, the available SAR data show that a large number of 370

modifications of the lipophilic tails are tolerated. With respect to selectivity, our data do not 371

provide any obvious structural or molecular explanation to rationalize the relative selectivity of 372

QN254 for P. falciparum versus the human enzyme. 373

For more than 30 years, it has been known that QN derivatives inhibit the growth and 374

proliferation of eukaryotic parasites such as Toxoplasma, Pneumocystis, Trypanosoma and 375

Plasmodium (8, 10, 11, 23, 36, 38, 45, 46). In fact, an extensive series of quinazolines have 376

previously proved to be efficacious in rodent and simian malaria models (41-44). The most 377

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

20

advanced candidate of this series, WR158122, entered clinical development as a candidate 378

antimalarial drug but was later abandoned because of its relatively poor oral efficacy in humans 379

(43). 380

Following oral administration QN254 is well absorbed, displays a long half-life and an 381

excellent oral bioavailability in rodents as well as in larger species such as dog (data not 382

shown). Although PAMPA and Caco-2 in vitro permeability data were predictive of the good 383

absorption and bioavailability, our in vitro metabolic stability data predicted that QN254 would 384

be subject to rapid hepatic clearance in all preclinical species as well as in humans (data not 385

shown). Nonetheless, in vivo QN254 showed a low (rats and dogs) to moderate (mouse) 386

systemic clearance in all preclinical species. We cannot explain the discrepancy between in 387

vitro and in vivo hepatic clearance and further studies would be required to determine the main 388

clearance mechanism in vivo. 389

Consistent with our pharmacokinetics results, QN254 displays good efficacy in the P. berghei 390

mouse model through the oral route. However, in P. berghei infected mice, we have found 391

evidence of toxicity upon repeated daily dosing for 3 days at doses as low as 60 mg/kg and 392

indeed further rat toxicology studies showed that QN254 does not possess an adequate 393

therapeutic window. At this juncture we cannot unambiguously establish whether the observed 394

toxicity is due to on- or off-target effects; and as an example we indeed have evidence that 395

QN254 presents some inherent risk of cardiotoxicity because of its ability to bind and inhibit 396

the hERG channel in vitro (hERG IC50 = 1.4 µM). Nonetheless, the rat histopathological data 397

showing that a highly proliferative tissue such as the bone marrow is massively affected is 398

consistent with on-target toxicity and sustained inhibition of folates metabolism through DHFR 399

inhibition (50, 54). Thus, despite its relative selectivity toward plasmodium DHFR enzyme in 400

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

21

vitro, QN254 appears to potently inhibit mammalian (rat) DHFR in vivo. Based on the 401

exposure reached at the efficacious dose and the exposure reached in rats upon dosing at 50 402

mg/kg in the toxicology study, we estimated that the therapeutic window is less than two. A 403

narrow therapeutic window is indeed a hallmark of almost all antifolates and has historically 404

been a significant hurdle to the development of new therapies based on DHFR inhibition. PM, 405

for example, displays an acute LD50 of 130 mg/kg in mice (50). Mitigating this toxicity risk is 406

the fact that antifolates such as PM are commonly used in combination at low doses (<5 mg/kg 407

in humans). Because of the severe toxicity observed in rats, the Novartis Institute for Tropical 408

Diseases is no longer pursuing the development of QN254. Nevertheless, further research into 409

the feasibility of strategies—e.g. pro-drug approach, supplementation with folate derivatives 410

such as 5-methyl-THF (27), or combination with antimalarials showing synergistic effects with 411

DHFR inhibitors—aiming to increase the therapeutic window of QN254 may be warranted. 412

413

6 Acknowledgments 414

We thank the director of the Kenya Medical Research Institute for permission to publish these 415

data. SK is an international research scholar of Howard Hughes Medical Institute (HHMI), 416

USA. KEMRI received support from the EU Commission under Framework 6 as part of the 417

AntiMal Integrated Project 018834, the European & Developing Countries Clinical Trials 418

Partnership (EDCTP) and the Wellcome Trust WT077092. BIOTEC was supported by grants 419

from Thailand-TDR (T-2) and the Medicines for Malaria Venture. NITD and STI receive 420

funding from a joint grant of the Medicines for Malaria Venture and the Wellcome Trust 421

(WT078285). 422

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

22

423

7 References 424

1. Brunger, A. T., P. D. Adams, G. M. Clore, W. L. DeLano, P. Gros, R. W. Grosse-Kunstleve, J. S. 425

Jiang, J. Kuszewski, M. Nilges, N. S. Pannu, R. J. Read, L. M. Rice, T. Simonson, and G. L. 426

Warren. 1998. Crystallography & NMR system: A new software suite for macromolecular structure 427

determination. Acta Crystallogr. D 54 ( Pt 5):905-921. 428

2. Canfield, C. J., W. K. Milhous, A. L. Ager, R. N. Rossan, T. R. Sweeney, N. J. Lewis, and D. P. 429

Jacobus. 1993. PS-15: a potent, orally active antimalarial from a new class of folic acid antagonists. Am. 430

J. Trop. Med. Hyg. 49:121-126. 431

3. Childs, G. E., and C. Lambros. 1986. Analogues of N-benzyloxydihydrotriazines: in vitro antimalarial 432

activity against Plasmodium falciparum. Ann. Trop. Med. Parasitol. 80:177-181. 433

4. Chitnumsub, P., J. Yuvaniyama, J. Vanichtanankul, S. Kamchonwongpaisan, M. D. Walkinshaw, 434

and Y. Yuthavong. 2004. Characterization, crystallization and preliminary X-ray analysis of 435

bifunctional dihydrofolate reductase-thymidylate synthase from Plasmodium falciparum. Acta 436

Crystallogr. D Biol. Crystallogr. 60:780-783. 437

5. Cody, V., N. Galitsky, J. R. Luft, W. Pangborn, S. F. Queener, and A. Gangjee. 2002. Analysis of 438

quinazoline and pyrido[2,3-d]pyrimidine N9-C10 reversed-bridge antifolates in complex with NADP+ 439

and Pneumocystis carinii dihydrofolate reductase. Acta Crystallogr D Biol Crystallogr 58:1393-1399. 440

6. Delano W L. 2002. The PyMOL Molecular Graphics System. The PyMOL Molecular Graphics 441

System.:Delano Scientific, Palo Alto, CA. 442

7. Dondorp, A. M., F. Nosten, P. Yi, D. Das, A. P. Phyo, J. Tarning, K. M. Lwin, F. Ariey, W. 443

Hanpithakpong, S. J. Lee, P. Ringwald, K. Silamut, M. Imwong, K. Chotivanich, P. Lim, T. 444

Herdman, S. S. An, S. Yeung, P. Singhasivanon, N. P. Day, N. Lindegardh, D. Socheat, and N. J. 445

White. 2009. Artemisinin resistance in Plasmodium falciparum malaria. N. Engl. J. Med. 361:455-467. 446

8. Elslager, E. F., O. D. Bird, J. Clarke, S. C. Perricone, and D. F. Worth. 1972. Folate antagonists. 9. 447

2,4-Diamino-6-((aralkyl)alkylamino)quinazolines, a potent class of antimetabolites with prodigious 448

antimalarial effects. J. Med. Chem. 15:1138-1146. 449

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

23

9. Elslager, E. F., J. Clarke, L. M. Werbel, D. F. Worth, and J. Davoll. 1972. Folate antagonists. 3. 2,4-450

Diamino-6-(heterocyclic)quinazolines, a novel class of antimetabolites with potent antimalarial and 451

antibacterial activity. J. Med. Chem. 15:827-836. 452

10. Elslager, E. F., M. P. Hutt, P. Jacob, J. Johnson, B. Temporelli, L. M. Werbel, D. F. Worth, and L. 453

Rane. 1979. Folate antagonists. 15. 2,3-Diamino-6-(2-naphthylsulfonyl)quinazoline and related 2,4-454

diamino-6-[(phenyl and naphthyl)sulfinyl and sulfonyl]quinazolines, a potent new class of 455

antimetabolites with phenomenal antimalarial activity. J. Med. Chem. 22:1247-1257. 456

11. Elslager, E. F., P. Jacob, J. Johnson, L. M. Werbel, D. F. Worth, and L. Rane. 1978. Folate 457

antagonists. 13. 2,4-Diamino-6-](alpha,alpha,alpha-trifluoro-m-tolyl)thio]quinazoline and related 2,4-458

diamino-6-[(phenyl- and naphthyl)thio]quinazolines, a unique class of antimetabolites with extraordinary 459

antimalarial and antibacterial effects. J. Med. Chem. 21:1059-1070. 460

12. Fidock, D. A., P. J. Rosenthal, S. L. Croft, R. Brun, and S. Nwaka. 2004. Antimalarial drug 461

discovery: efficacy models for compound screening. Nat. Rev. Drug Discov. 3:509-520. 462

13. Gangjee, A., S. Kurup, and O. Namjoshi. 2007. Dihydrofolate reductase as a target for chemotherapy 463

in parasites. Curr. Pharm. Des. 13:609-639. 464

14. Gregson, A., and C. V. Plowe. 2005. Mechanisms of resistance of malaria parasites to antifolates. 465

Pharmacol. Rev. 57:117-145. 466

15. Jackman, A. L., F. T. Boyle, and K. R. Harrap. 1996. Tomudex (ZD1694): from concept to care, a 467

programme in rational drug discovery. Invest. New Drugs 14:305-316. 468

16. Jones, T. A., J. Y. Zou, S. W. Cowan, and M. Kjeldgaard. 1991. Improved methods for building 469

protein models in electron density maps and the location of errors in these models. Acta Crystallogr. A 470

47:110-119. 471

17. Kamchonwongpaisan, S., R. Quarrell, N. Charoensetakul, R. Ponsinet, T. Vilaivan, J. 472

Vanichtanankul, B. Tarnchompoo, W. Sirawaraporn, G. Lowe, and Y. Yuthavong. 2004. Inhibitors 473

of multiple mutants of Plasmodium falciparum dihydrofolate reductase and their antimalarial activities. J 474

Med. Chem. 47:673-680. 475

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

24

18. Kamchonwongpaisan, S., J. Vanichtanankul, B. Tarnchompoo, J. Yuvaniyama, S. Taweechai, and 476

Y. Yuthavong. 2005. Stoichiometric selection of tight-binding inhibitors by wild-type and mutant forms 477

of malarial (Plasmodium falciparum) dihydrofolate reductase. Anal. Chem. 77:1222-1227. 478

19. Kiara, S. M., J. Okombo, V. Masseno, L. Mwai, I. Ochola, S. Borrmann, and A. Nzila. 2009. In vitro 479

activity of antifolate and polymorphism in dihydrofolate reductase of Plasmodium falciparum isolates 480

from Kenyan coast: Emergence of parasites with Ile-164-Leu mutation. Antimicrob. Agents Chemother. 481

20. Kinyanjui, S. M., E. K. Mberu, P. A. Winstanley, D. P. Jacobus, and W. M. Watkins. 1999. The 482

antimalarial triazine WR99210 and the prodrug PS-15: folate reversal of in vitro activity against 483

Plasmodium falciparum and a non-antifolate mode of action of the prodrug. Am. J. Trop. Med. Hyg. 484

60:943-947. 485

21. Kleywegt G J. 1997. Validation of protein models from CA coordinates alone. J. Mol. Biol. 273:371-486

376. 487

22. Knight, D. J., P. Mamalis, and W. Peters. 1982. The antimalarial activity of N-benzyl-488

oxydihydrotriazines. III. The activity of 4,6-diamino-1,2-dihydro-2,2-dimethyl-1-(2,4,5,-489

trichloropropyloxy)-1,3,5-t riazine hydrobromide (BRL 51084) and hydrochloride (BRL 6231). Ann 490

Trop. Med. Parasitol. 76:1-7. 491

23. Kovacs, J. A., C. J. Allegra, B. A. Chabner, J. C. Swan, J. Drake, M. Lunde, J. E. Parrillo, and H. 492

Masur. 1987. Potent effect of trimetrexate, a lipid-soluble antifolate, on Toxoplasma gondii. J. Infect. 493

Dis. 155:1027-1032. 494

24. Laskowski R A, MacArthur M W, Moss D S, and Thronton J M. 1993. PROCHECK: a program to 495

check the stereochemical quality of protein structures. . J. Appl. Cryst. 26:283-291. 496

25. Morris A L, MacArthur M W, Hutchinson E G, and Thronton J M. 1992. Stereochemical quality of 497

protein structure coordinates. . Proteins. 12:345-364. 498

26. Murshudov G N, Vagin A A, and Dodson E J. 1997. Refinement of macromolecular structures by the 499

maximum-likelihood method. Acta Crystallogr. D 53:240-255. 500

27. Nduati, E., A. Diriye, S. Ommeh, L. Mwai, S. Kiara, V. Masseno, G. Kokwaro, and A. Nzila. 2008. 501

Effect of folate derivatives on the activity of antifolate drugs used against malaria and cancer. Parasitol. 502

Res. 102:1227-1234. 503

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

25

28. Noedl, H., Y. Se, K. Schaecher, B. L. Smith, D. Socheat, and M. M. Fukuda. 2008. Evidence of 504

artemisinin-resistant malaria in western Cambodia. N. Engl. J. Med. 359:2619-2620. 505

29. Nzila, A. 2006. Inhibitors of de novo folate enzymes in Plasmodium falciparum. Drug Discov. Today 506

11:939-944. 507

30. Nzila, A. 2006. The past, present and future of antifolates in the treatment of Plasmodium falciparum 508

infection. J Antimicrob. Chemother. 57:1043-1054. 509

31. Nzila, A. M., E. K. Mberu, J. Sulo, H. Dayo, P. A. Winstanley, C. H. Sibley, and W. M. Watkins. 510

2000. Towards an understanding of the mechanism of pyrimethamine-sulfadoxine resistance in 511

Plasmodium falciparum: genotyping of dihydrofolate reductase and dihydropteroate synthase of Kenyan 512

parasites. Antimicrob. Agents Chemother. 44:991-996. 513

32. O'Dwyer, P. J., R. J. DeLap, S. A. King, A. J. Grillo-Lopez, D. F. Hoth, and B. Leyland-Jones. 514

1987. Trimetrexate: clinical development of a nonclassical antifolate. NCI Monogr:105-109. 515

33. O'Dwyer, P. J., D. D. Shoemaker, J. Plowman, J. Cradock, A. Grillo-Lopez, and B. Leyland-Jones. 516

1985. Trimetrexate: a new antifol entering clinical trials. Invest. New Drugs 3:71-75. 517

34. Ommeh, S., E. Nduati, E. Mberu, G. Kokwaro, K. Marsh, A. Rosowsky, and A. Nzila. 2004. In vitro 518

activities of 2,4-diaminoquinazoline and 2,4-diaminopteridine derivatives against Plasmodium 519

falciparum. Antimicrob. Agents Chemother. 48:3711-3714. 520

35. Otwinowski, Z. M., and W. Minor. 1979. Processing of X-ray diffraction data collected in oscillation 521

mode. Method Enzymol. 276:307-326. 522

36. Piper, J. R., C. A. Johnson, C. A. Krauth, R. L. Carter, C. A. Hosmer, S. F. Queener, S. E. Borotz, 523

and E. R. Pfefferkorn. 1996. Lipophilic antifolates as agents against opportunistic infections. 1. Agents 524

superior to trimetrexate and piritrexim against Toxoplasma gondii and Pneumocystis carinii in in vitro 525

evaluations. J. Med. Chem. 39:1271-1280. 526

37. Ridley, R. G., H. Matile, C. Jaquet, A. Dorn, W. Hofheinz, W. Leupin, R. Masciadri, F. P. Theil, 527

W. F. Richter, M. A. Girometta, A. Guenzi, H. Urwyler, E. Gocke, J. M. Potthast, M. Csato, A. 528

Thomas, and W. Peters. 1997. Antimalarial activity of the bisquinoline trans-N1,N2-bis (7-529

chloroquinolin-4-yl)cyclohexane-1,2-diamine: comparison of two stereoisomers and detailed evaluation 530

of the S,S enantiomer, Ro 47-7737. Antimicrob. Agents Chemother. 41:677-686. 531

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

26

38. Rosowsky, A., C. E. Mota, J. E. Wright, J. H. Freisheim, J. J. Heusner, J. J. McCormack, and S. F. 532

Queener. 1993. 2,4-Diaminothieno[2,3-d]pyrimidine analogues of trimetrexate and piritrexim as 533

potential inhibitors of Pneumocystis carinii and Toxoplasma gondii dihydrofolate reductase. J. Med. 534

Chem. 36:3103-3112. 535

39. Rosowsky, A., C. E. Mota, J. E. Wright, and S. F. Queener. 1994. 2,4-Diamino-5-chloroquinazoline 536

analogues of trimetrexate and piritrexim: synthesis and antifolate activity. J. Med. Chem. 37:4522-4528. 537

40. Sasi, P., A. Abdulrahaman, L. Mwai, S. Muriithi, J. Straimer, E. Schieck, A. Rippert, M. 538

Bashraheil, A. Salim, J. Peshu, K. Awuondo, B. Lowe, M. Pirmohamed, P. Winstanley, S. Ward, A. 539

Nzila, and S. Borrmann. 2009. In vivo and in vitro efficacy of amodiaquine against Plasmodium 540

falciparum in an area of continued use of 4-aminoquinolines in East Africa. J. Infect. Dis. 199:1575-541

1582. 542

41. Schmidt, L. H. 1978. Plasmodium falciparum and Plasmodium vivax infections in the owl monkey 543

(Aotus trivirgatus). III. Methods employed in the search for new blood schizonticidal drugs. Am. J. Trop. 544

Med. Hyg. 27:718-737. 545

42. Schmidt, L. H. 1979. Studies on the 2,4-diamino-6-substituted quinazolines. II. Activities of selected 546

derivatives against infections with various drug-susceptible and drug-resistant strains of Plasmodium 547

falciparum and Plasmodium vivax in owl monkeys. Am. J. Trop. Med. Hyg. 28:793-807. 548

43. Schmidt, L. H. 1979. Studies on the 2,4-diamino-6-substituted quinazolines. III. The capacity of 549

sulfadiazine to enhance the activities of WR-158,122 and WR-159,412 against infections with various 550

drug-susceptible and drug-resistant strains of Plasmodium falciparum and Plasmodium vivax in owl 551

monkeys. Am. J. Trop. Med. Hyg. 28:808-818. 552

44. Schmidt, L. H., and R. N. Rossan. 1979. Studies on the 2,4-diamino-6 substituted quinazolines. I. 553

Antimalarial activities of 2,4-diamino-6-[(3,4-dichlorobenzyl)-nitrosoamino]-quinazoline (CI-679) as 554

exhibited in rhesus monkeys infected with the Ro or Ro/PM strains of Plasmodium cynomolgi. Am. J. 555

Trop. Med. Hyg. 28:781-792. 556

45. Senkovich, O., V. Bhatia, N. Garg, and D. Chattopadhyay. 2005. Lipophilic antifolate trimetrexate is 557

a potent inhibitor of Trypanosoma cruzi: prospect for chemotherapy of Chagas' disease. Antimicrob. 558

Agents Chemother. 49:3234-3238. 559

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

27

46. Senkovich, O., B. Pal, N. Schormann, and D. Chattopadhyay. 2003. Trypanosoma cruzi genome 560

encodes a pteridine reductase 2 protein. Mol. Biochem .Parasitol. 127:89-92. 561

47. Sibley, C. H., J. E. Hyde, P. F. Sims, C. V. Plowe, J. G. Kublin, E. K. Mberu, A. F. Cowman, P. A. 562

Winstanley, W. M. Watkins, and A. M. Nzila. 2001. Pyrimethamine-sulfadoxine resistance in 563

Plasmodium falciparum: what next? Trends Parasitol. 17:582-588. 564

48. Sirawaraporn, W., R. Sirawaraporn, S. Yongkiettrakul, A. Anuwatwora, G. Rastelli, S. 565

Kamchonwongpaisan, and Y. Yuthavong. 2002. Mutational analysis of Plasmodium falciparum 566

dihydrofolate reductase: the role of aspartate 54 and phenylalanine 223 on catalytic activity and antifolate 567

binding. Mol. Biochem. Parasitol. 121:185-193. 568

49. Sixsmith, D. G., W. M. Watkins, J. D. Chulay, and H. C. Spencer. 1984. In vitro antimalarial activity 569

of tetrahydrofolate dehydrogenase inhibitors. Am. J. Trop. Med. Hyg. 33:772-776. 570

50. Stephan-Guldner, M. 1993. Preclinical toxicology and safety pharmacology of brodimoprim in 571

comparison to trimethoprim and analogs. J. Chemother. 5:400-410. 572

51. Tarnchompoo, B., C. Sirichaiwat, W. Phupong, C. Intaraudom, W. Sirawaraporn, S. 573

Kamchonwongpaisan, J. Vanichtanankul, Y. Thebtaranonth, and Y. Yuthavong. 2002. 574

Development of 2,4-diaminopyrimidines as antimalarials based on inhibition of the S108N and 575

C59R+S108N mutants of dihydrofolate reductase from pyrimethamine-resistant Plasmodium falciparum. 576

J. Med. Chem. 45:1244-1252. 577

52. Vennerstrom, J. L., S. Arbe-Barnes, R. Brun, S. A. Charman, F. C. Chiu, J. Chollet, Y. Dong, A. 578

Dorn, D. Hunziker, H. Matile, K. McIntosh, M. Padmanilayam, J. Santo Tomas, C. Scheurer, B. 579

Scorneaux, Y. Tang, H. Urwyler, S. Wittlin, and W. N. Charman. 2004. Identification of an 580

antimalarial synthetic trioxolane drug development candidate. Nature 430:900-904. 581

53. White, N. J. 2008. Qinghaosu (artemisinin): the price of success. Science 320:330-334. 582

54. Winstanley, P. A., E. K. Mberu, I. S. Szwandt, A. M. Breckenridge, and W. M. Watkins. 1995. In 583

vitro activities of novel antifolate drug combinations against Plasmodium falciparum and human 584

granulocyte CFUs. Antimicrob. Agents Chemother. 39:948-952. 585

55. Yuthavong, Y., J. Yuvaniyama, P. Chitnumsub, J. Vanichtanankul, S. Chusacultanachai, B. 586

Tarnchompoo, T. Vilaivan, and S. Kamchonwongpaisan. 2005. Malarial (Plasmodium falciparum) 587

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

28

dihydrofolate reductase-thymidylate synthase: structural basis for antifolate resistance and development 588

of effective inhibitors. Parasitol. 130:249-259. 589

56. Yuvaniyama, J., P. Chitnumsub, S. Kamchonwongpaisan, J. Vanichtanankul, W. Sirawaraporn, P. 590

Taylor, M. D. Walkinshaw, and Y. Yuthavong. 2003. Insights into antifolate resistance from malarial 591

DHFR-TS structures. Nat. Struct. Biol. 10:357-365. 592

593

594

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

29

Figure legends 595

596

Figure 1. Co-crystal structure of dihydrofolate reductase quadruple mutant (mutations at 597

codons 108, 51, 59, 164) showing H-bond interactions with QN254 (A) and WR99210 (B). 598

Both complexes are superposed in (C). 599

600

601

Figure 2. Pharmacokinetic profile of QN254 upon intravenous and oral administration – 602

QN254 was administered orally at 25 mg/kg (filled diamonds) and intravenously at 5 mg/kg 603

(open squares) to mice (A) and rats (B). Plasma concentrations were measured over time and 604

are indicated in ng/ml ± standard deviations (n=3). Note that the timescale is different in A and 605

B. Drug levels for the last time points in profile A were detectable but below the lower limit of 606

quantification (12 ng/ml). 607

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

30

608

Table legends 609

610

Table 1. Results of competitive binding kinetics experiments on DHFR enzymes of P. 611

falciparum wild-type (WT-PfDHFR) and quadruple mutant (QM-PfDHFR) [S108N, N51I, 612

C59R, and I164L], and human (hDHFR) against QN254, WR99210, pyrimethamine (PM) and 613

cycloguanil (CLG). Ki, binding constant (nM) 614

615

616

Table 2. Relationship between pyrimethamine (PM) and QN254 median inhibitory 617

concentrations that kill 50% of parasitemia (IC50) and dihydrofolate reductase (dhfr) genotypes 618

of clinical isolates. No wild-type dhfr isolate was identified; n, number of isolates. Double 619

mutants are S108N and N51I or S108N and C59R; triple mutants are S108, N51I and C59R; 620

and quadruple mutants are S108N, N51, C59R and Ile164Leu 621

622

623

Table 3. QN254 antimalarial activity in the Plasmodium. berghei murine malaria model. 624

Results represent % inhibition of parasitemia compared to untreated controls. QN254 was 625

prepared in 0.5% carboxymethyl cellulose; Survival of control animals: 5.6–6.2 days; * Mice 626

died around day 10, and were parasite-free; # Solution formulation with 10% ethanol, 30% 627

PEG400 (polyethylene glycol 400) and 60% of a 10% Vitamin ETPGS solution. N/A not 628

available; n≥10 mice except for the three times treatment with QN254 629

630

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

Table 1:

Ki (nM) Ki ratio

WT-PfDHFR

QM-

PfDHFR

hDHFR hDHFR/WT-

pfDHFR QM-

PfDHFR/WT-

PfDHFR

CLG 1.51 ± 0.1 454 ± 38 55.6 ± 7.8 37 300

PM 0.59 ± 0.05 385 ± 163 30.8 ± 1.4 52 652

WR99210 0.5 ± 0.1 1.9 ± 0.8 7.7 ± 0.4 15 1.8

QN254 0.39 ± 0.05 0.58 ± 0.06 10.2 ± 0.6 26 1.5

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

(A)

(B)

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

(C)

Figure 1:

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

Figure 2.

1

10

100

1000

10000

0 5 10 15 20 25

Time (h)

Mean C

oncentr

ation

(ng

/ml)

1

10

100

1000

10000

0 5 10 15 20 25

Time (h)

Mean C

oncentr

ation

(ng

/ml)

A

B

1

10

100

1000

10000

0 10 20 30 40 50

Time (h)

Mean C

oncentr

ation

(ng

/ml)

1

10

100

1000

10000

0 10 20 30 40 50

Time (h)

Mean C

oncentr

ation

(ng

/ml)

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

Table 2:

.

*Data pertaining to PM were already published elsewhere (38).

**None of the tested differences between double and triple mutants were significant at p=0.05 (Kruskall–

Wallis test).

DHFR genotype IC50 (n)

nM

PM* QN254

Double mutant** 371 (8) 4.48 (6)

Triple mutant** 778.28 (24) 11.66 (20)

Quadruple mutant 3690.79 (1) 7.61 (1)

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from

Table 3:

Dosing regimen Parasitemia

reduction (%)

Average mouse

survival (days)

Cure rate

(%)

Toxicity rate

(%)

QN254

1 x 10 mg/kg 59 6.0 0 0

1 x 15 mg/kg 73 6.0 0 0

1 x 30 mg/kg >99.99 8.1 0 0

1 x 60 mg/kg >99.99 12.0 0 0

1 x 100 mg/kg >99.99 23.4 50 0

3 x 30 mg/kg >99.99 28.4 80 0

3 x 60 mg/kg >99.99 N/A 60 40*

3 x 100 mg/kg >99.99 N/A 20 80*

Artesunate#

1 x 100 mg/kg 98 7.0 0 0

3 x 30 mg/kg 98.7 12.2 0 0

Chloroquine#

1 x 100 mg/kg 99.6 12.5 0 0

3 x 30 mg/kg 99.8 14.3 0 0

Mefloquine#

1 x 100 mg/kg 89 25.5 0 0

3 x 30 mg/kg 98.8 22.1 0 0

Results represent % inhibition of parasitemia compared to untreated controls; QN254 was

prepared in 0.5% carboxymethyl cellulose; Survival of control animals: 5.6–6.2 days; * Mice

died around day 10, and were parasite-free; #や Solution formulation with 10% ethanol, 30%

PEG400 (polyethylene glycol 400) and 60% of a 10% Vitamin ETPGS solution. N/A not

available; nœ10 mice except for the three times treatment with QN254 because of ethical reasons

due to the observed toxicity

on June 9, 2018 by guesthttp://aac.asm

.org/D

ownloaded from


Recommended