Next Article in Journal
Structural and Thermodynamic Analysis of the Resistance Development to Pimodivir (VX-787), the Clinical Inhibitor of Cap Binding to PB2 Subunit of Influenza A Polymerase
Next Article in Special Issue
Identification of Kaurane-Type Diterpenes as Inhibitors of Leishmania Pteridine Reductase I
Previous Article in Journal
Sustainable Drying and Torrefaction Processes of Miscanthus for Use as a Pelletized Solid Biofuel and Biocarbon-Carrier for Fertilizers
Previous Article in Special Issue
The Alkaloid-Enriched Fraction of Pachysandra terminalis (Buxaceae) Shows Prominent Activity against Trypanosoma brucei rhodesiense
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nopol-Based Quinoline Derivatives as Antiplasmodial Agents

by
Rogers J. Nyamwihura
,
Huaisheng Zhang
,
Jasmine T. Collins
,
Olamide Crown
and
Ifedayo Victor Ogungbe
*
Department of Chemistry, Physics, and Atmospheric Sciences, Jackson State University, Jackson, MS 39217, USA
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(4), 1008; https://doi.org/10.3390/molecules26041008
Submission received: 17 January 2021 / Revised: 8 February 2021 / Accepted: 11 February 2021 / Published: 14 February 2021

Abstract

:
Malaria remains a significant cause of morbidity and mortality in Sub-Saharan Africa and South Asia. While clinical antimalarials are efficacious when administered according to local guidelines, resistance to every class of antimalarials is a persistent problem. There is a constant need for new antimalarial therapeutics that complement parasite control strategies to combat malaria, especially in the tropics. In this work, nopol-based quinoline derivatives were investigated for their inhibitory activity against Plasmodium falciparum, one of the parasites that cause malaria. The nopyl-quinolin-8-yl amides (24) were moderately active against the asexual blood stage of chloroquine-sensitive strain Pf3D7 but inactive against chloroquine-resistant strains PfK1 and PfNF54. The nopyl-quinolin-4-yl amides and nopyl-quinolin-4-yl-acetates analogs were generally less active on all three strains. Interesting, the presence of a chloro substituent at C7 of the quinoline ring of amide 8 resulted in sub-micromolar EC50 in the PfK1 strain. However, 8 was more than two orders of magnitude less active against Pf3D7 and PfNF54. Overall, the nopyl-quinolin-8-yl amides appear to share similar antimalarial profile (asexual blood-stage) with previously reported 8-aminoquinolines like primaquine. Future work will focus on investigating the moderately active and selective nopyl-quinolin-8-yl amides on the gametocyte or liver stages of Plasmodium falciparum and Plasmodium vivax.

Graphical Abstract

1. Introduction

Malaria is endemic in 91 countries, which accounts for 40% of the world’s population. According to the US Centers for Disease Control and Prevention, an estimated 228 million malaria cases occurred worldwide in 2018, and 405,000 people died [1]. It remains a serious public health problem in many developing tropical countries. There is no approved vaccine to prevent malaria for the general population. However, there are several therapies whose effectiveness is limited by the strain of Plasmodium responsible for the disease or by drug resistance. Artemisinin-based combination therapies (ACTs) are the first-line drugs for the treatment of malaria caused by P. falciparum in most malaria-endemic countries, while chloroquine (CQ) or ACT is used for the treatment of uncomplicated P. vivax in countries endemic for vivax malaria. Combinations of primaquine with CQ or ACTs are used for radical cure of relapsing malaria parasites due to P. vivax and P. ovale [2]. The 8-aminoquinoline-based drug primaquine (PQ) is also used as a gametocytocidal agent to reduce the transmission of P. falciparum in areas with verified artemisinin resistance [3]. 4-aminoquinoline-based compounds like chloroquine are known for their ability to accumulate as dicationic molecules in the plasmodial food vacuole and inhibit heme detoxification [4,5,6]. Since malarial parasites lack the oxygenase required to metabolize heme, soluble heme (iron-protoporphyrin IX) is converted into an insoluble and inert form called hemozoin, which the parasite can efficiently excrete. When these quinoline-based antimalarial drugs interfere with the conversion of soluble heme to hemozoin, heme accumulates and becomes toxic to the parasite because of radical reactions, damaged cell membranes, and cellular proteopathy [7,8]. The development of resistance to 4-aminoquinoline-based compounds such as CQ is linked to mutations in CQ resistance transporter (PfCRT) in P. falciparum and upregulation in the expression of CQ resistance transporter (PvCRT) in P. vivax. Mutations in P. falciparum multidrug-resistant protein 1 have also been linked to resistance to artemisinin, halofantrine, mefloquine, and quinine. The mutations and/or upregulated expression shield the parasites from drug accumulation and are viewed as consequences of evolutionary pressure to increase the rate of removal of drugs from the compartment containing the putative drug target(s) [9,10,11,12]. The mechanism of action of 8-aminoquinoline-based antimalarials is unclear. However, some studies have suggested that it involves the generation of reactive oxygen species as well as inhibition of the electron transport chain because of metabolic transformation [13,14]. Quinoline-based compounds remain a robust class of molecules for new antimalarial agents. In this study, quinoline compounds that possess an α-pinene scaffold were synthesized (Scheme 1) and evaluated for their antiplasmodial activities. α-pinene is a common constituent of essential oils, and it was previously shown to have antiplasmodial activity [15]. Also, phloroglucinol-terpenes robustadial A and B from Eucalyptus robusta, containing an α-pinene moiety, have been reported to have antiplasmodial activities [16].

2. Results and Discussion

The synthetic route to the compounds is shown in Scheme 1 below. To obtain amides 16 and esters 712, (1R)-(−)-nopol (N1) was oxidized to its corresponding acid using Jones reagent, and the acid was coupled with the corresponding aminoquinolines and quinolinols. Oxidation of nopol produced a relatively low yield of nopoic acid (N2). This is likely due to cationic rearrangement of the pinene ring to relieve angle strain caused by the four-membered ring bridgehead [17]. The initial reaction of nopoic acid with 5-methoxy-8-aminoquinoline and 8-aminoquinoline, using HBTU-HOBT as coupling reagents, produced two stereoisomers each (14). The diastereotopic α-protons in 1 (δ 3.32 and 3.13 ppm) and 3 (δ 3.34 and 3.16 ppm) share similar magnetic fields compared to the diastereotopic α-protons in 2 (δ 3.56–3.48 and 2.88 ppm) and 4 (δ 3.55–3.48 and 2.92–2.82 ppm). The observed anisotropic effect is likely due to two different orientations of the quinoline ring in the two sets of isomers. The amides (14) were subsequently tested for their antiplasmodial activities in vitro using the P. falciparum 3D7 (Pf3D7, chloroquine-sensitive strain), and all four compounds showed moderate but selective low micromolar potency against the parasite (Table 1). The bioactivity (EC50) of the compounds was similar to reported EC50 values for 8-aminoquinolines like primaquine [18]. This was particularly interesting because the compounds lack the essential side-chain amino moiety typical of quinoline-based antimalarials [19]. Therefore, to extend the compound’s series and to investigate the potentially more potent 4-aminoquinoline derivatives, a set of 4-aminoquinoline-based analogs (512) were synthesized and evaluated against Pf3D7. Amides 5 and 6 were weakly active (EC50 26.4 and 17.1 µM, respectively) against Pf3D7, and 5 was significantly more cytotoxic to HepG2 (hepatocarcinoma) cells compared to the corresponding 8-aminoquinolyl analogs (3 and 4, Table 1). Esters 7–12 were generally inactive against Pf3D7 apart from 9 (EC50 6.2 µM). Three other esters, 1315, were synthesized using 7-chloro-2-methyl-4-quinolinecarboxylic acid, 2-chloro-4-quinolinecarboxylic acid, and 2-indolecarboxylic acid, respectively. Compound 13 displayed very weak antiplasmodial activity while 14 and 15 were moderately active against Pf3D7. The 4-aminoquinoline nopyl amides and esters displayed much weaker activities against Pf3D7 than the 8-aminoquinoline-derived compounds. To determine if the compounds are active against strains of P. falciparum that are less sensitive to chloroquine, compounds 38 and 14 were assayed against the NF54 and K1 strains of P. falciparum. Compounds 3, 4, and 14 were inactive against both strains, while 5 and 6 were weakly active against both strains (Table 2). Surprisingly, compound 8, which bears the 7-chloroquinolyl moiety, was significantly more active against the “multidrug-resistant” K1 strain than the NF54 and 3D7 strains (EC50 PfK1 0.16 µM vs. EC50 NF54 23.3 µM and EC50 Pf3D7 > 50 µM). It is unclear why the K1 was more susceptible to 8. However, a previous report has shown that mutations in CQ resistance transporter, like in the K1 strain, can enhance antimalarial agents’ antiplasmodial activities against CQ-resistant strains [20]. For example, the antiviral agent amantadine can accumulate in the cytosol, act on its molecular target, and exert more significant antiplasmodial activity because the 76I-PfCRTK1 transporter variant can transport it back into the cytosol from the food vacuole more effectively than the wild-type transporter (PfCRTK1). It is likely that 8 is transported with greater efficiency by PfCRT in the K1 strain and able to engage effectively with a putative cytosolic target, or its transportation out of the food vacuole is significantly hindered in the K1 strain and therefore displayed a more substantial anti-hemozoin effect in the parasite.
Synthetic efforts to introduce epoxy and vicinal diol groups to increase aqueous solubility via the unsaturation in the nopol substructure for 115 were unsuccessful [21,22,23]. However, 16 was synthesized via epoxidation of nopol followed by esterification reaction with 7-chloro-2-methyl-4-quinolinecarboxylic acid using EDCI and DMAP. Compound 16 was weakly active but nonselective against Pf3D7.
In conclusion, the 8-aminoquinolyl amides of nopol were found to be moderately active against the chloroquine-sensitive strain of P. falciparum (Pf3D7), while the corresponding 4-aminoquinolyl analogs were generally more active on the multidrug-resistant K1 strain, especially ester 8. Future efforts will be directed towards (1) evaluating the antiplasmodial activity of the 8-aminoquinolyl amides on different life stages of P. falciparum, as well as on P. vivax and P. ovale, and (2) adding an amino functionality to the α-pinene or another similar scaffold to potentially enhance the antiplasmodial activities and to understand further the structure–activity relationships of compounds 14, 8, 9, and 14.

3. Materials and Methods

3.1. Compound Characterization

All reagents were purchased from commercial sources and used without further purification. The 1H and 13C-NMR data were obtained Varian 500 MHz (Agilent, Santa Clara, CA, USA) and Bruker Ultrashield Avance 400 (Bruker, Billerica, MA, USA) spectrometers. Thin layer chromatography (TLC) and NMR were used to monitor reactions and determine compound purity. Compounds were purified by silica gel (Sortech, Norcross, GA, USA, 60 Å, 200–500 µm (35 × 70 mesh)) column chromatography or on pre-coated preparative TLC plates (SiliCycle Inc, Quebec City, Canada, 1000 μm, 20 × 20 cm). High-resolution mass spectrometric (HRMS) data was obtained on a Synapt G2 HDMS instrument operated in positive or negative ESI mode. Spectra are provided as supplementary information (Supplementary Figure S1.1a–S1.16b).

3.2. Synthesis of (R)-(−)-Nopoic Acid Using Jones Reagent

(R)-(−)-Nopol (2.2 g, Sigma-Aldrich, St. Louis, MO, USA, 98% purity) was placed in a round bottom flask containing a magnetic stirring bar, and 20 mL of acetone was added. The solution was stirred in an ice-water bath for 10 min. To the solution, 10 mL of 3.88 M Jones reagent (Made from Fisher Scientific’s Chromium(VI) oxide, 99+% purity)in acetone was slowly added for 1 h. TLC was used to monitor the reaction progress. Once no more nopol was detected on analytical TLC, 1–2 mL of isopropanol was added to consume excess Jones reagent. The reaction turned brownish with the subsequent formation of dark green precipitate at the bottom of the flask. The reaction was washed with 20 mL of saturated NaHCO3 solution, and the crude product was extracted with 45 mL of ethyl acetate (EtOAc). The EtOAc extract was washed with brine and dried over MgSO4 and concentrated under vacuum. Pure nopoic acid was obtained by column chromatography using 85:1 CHCl3/ethyl acetate (Rf = 0.32). Pure nopoic acid was obtained in 30% yield. 1H-NMR (500 MHz, CDCl3) δ 11.48 (s, 1H, COOH), 5.42 (s, 1H, H-3′), 3.01 (dd, J = 15.7, 15.2 Hz, 2H, H-10′), 2.38 (m, 1H, H-6′), 2.33–2.19 (m, 2H, H-4′), 2.14 (td, J = 5.6, 1.6 Hz, 1H, H-5′), 2.09 (m, 1H, H-1′), 1.28 (s, 3H, H-9′), 1.22 (d, J = 8.7 Hz, 1H, H-6′), 0.83 (s, 3H, H-8′).13C-NMR (126 MHz, CDCl3) δ 178.0 (C=O), 140.5 (C-2′), 121.3 (C-3′), 45.9 (C-10′), 42.4 (C-5′), 40.5 (C-1′), 38.1 (C-7′), 31.7 (C-6′), 31.5 (C-4′), 26.2 (C-9′), 20.9 (C-8′).

3.3. Synthesis of (2-(6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl)-N-(5-methoxyquinolin-8-yl Acetamide (1 and 2)

To a round bottom flask containing 3 mL of N,N-dimethylformamide (DMF, Fisher Scientific, >99.8%), nopoic acid (200 mg, 1.1 mmol, 1.0 eq.) was added, followed by 0.1 mL of N,N-diisopropylethylamine (DIEA, Sigma-Aldrich, 99.5% purity), N,N,N′,N′-Tetramethyl-O-(1H-benzotriazol-1-yl)uronium hexafluorophosphate (HBTU, Sigma-Aldrich, 98% purity, 400.3 mg, 0.1 mmol), and HOBT (1-Hydroxybenzotriazole hydrate, 98% purity, 144.74 mg, 1.1 mmol, 1.0 eq.) [24]. The reaction mixture was stirred at room temperature for 30 min, and 5-methoxy-8-aminoquinoline (186.3 mg, 1.1 mmol, and 1.0 eq, Sigma-Aldrich, 95% purity) was added to it. The reaction was allowed to run for 16 h. TLC results show some unreacted amine and activated nopoic acid with two prominent product spots. The compounds (1 and 2) were purified using preparative TLC (10:1 hexane/ethyl acetate, Rf = 0.55 (1) and 0.68 (2)), and the overall product yield was 35%.

3.3.1. Compound 1

Molecules 26 01008 i017
1H-NMR (500 MHz, CDCl3) δ 9.81 (s, 1H, NH), 8.79 (dd, J = 4.2, 1.7 Hz, 1H, H-2), 8.68 (d, J = 8.6 Hz, 1H, H-6), 8.56 (dd, J = 8.4 Hz, J = 1.7 Hz, 1H, H-4), 7.42 (dd, J = 8.4 Hz, J = 4.2 Hz, 1H, H-3), 6.83 (d, J = 8.6 Hz, 1H, H-7), 5.66 (m, 1H, H-3′), 3.98 (s, 3H, OMe), 3.32 (d, J = 14.2 Hz, 1H, H-10′), 3.13 (m, 1H, H-10′), 2.43–2.39 (m, 2H, H-6′, H-4′), 2.35–2.29 (m, 1H, H-1′), 2.22 (m, 1H, H-5′), 2.13 (m, 1H, H-4′), 1.48 (d, J = 8.7 Hz, 1H, H-6′), 1.26 (s, 3H, H-9′), 0.85 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 169.1 (C=O), 150.1 (C-5), 148.6 (C-2), 142.2 (C-2′), 139.2 (C-8), 131.1 (C-6), 128.1 (C-8a), 122.3 (C7), 120.6 (C-3), 120.4 (C-4a), 116.4 (C-3′), 104.3 (C-4), 55.8 (OMe), 46.9 (1′), 45.9 (C-10′), 40.4 (C-5′), 38.1 (C-6′), 31.9 (C-7′), 31.6 (C-4′), 26.1 (C-9′), 20.9 (C-8′). HRMS: [M + H]+: 337.1916 m/z calculated for C21H25N2O2, found 337.1910 m/z.

3.3.2. Compound 2

Molecules 26 01008 i018
1H-NMR (500 MHz, CDCl3) δ 9.49 (s, 1H, NH), 8.79 (dd, J = 4.3, 1.9 Hz, 1H, H-2), 8.76 (d, J = 8.6 Hz, 1H, H-6), 8.56 (dd, J = 8.5, 1.8 Hz, 1H, H-4), 7.43 (dd, J = 8.3, 4.2 Hz, 1H, H-3), 6.84 (d, J = 8.6 Hz, 1H, H-7), 5.76 (d, J = 2.2 Hz, 1H, H-3′), 3.98 (s, 3H, OMe), 3.52 (m, 1H, H-10′), 2.88 (m, 1H, H-10′), 2.54 (m, 1H, H-6′), 2.41 (m, 1H, H-4′), 2.09 (m, 1H, H-1′), 2.01-1.88 (m, 2H, H-5′, H4′), 1.46 (d, J = 10.0 Hz, 1H, H-6′), 1.29 (s, 3H, H-9′), 0.80 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 165.2 (C=O), 149.8 (C-5), 148.5 (C-2), 139.2 (C-2′), 131.3 (C-6), 128.7 (C-8), 120.7 (C-7), 120.6 (C-8a), 117.2 (C-4a), 116.2 (C-3′), 116.0 (C-3), 104.6 (C-4), 55.9 (OMe), 53.9 (C-1′), 41.1 (C-10′), 40.7 (C-5′), 27.4 (C-6′), 26.3 (C-9′), 23.9 (C-7′), 22.4 (C-4′), 22.30 (C-8′). HRMS: [M + H]+: 337.1916 m/z calculated for C21H25N2O2, found 337.1916 m/z.

3.4. Synthesis of 2-(6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl)-N-(quinolin-8-yl) Acetamide (3 and 4)

Compounds 3 and 4 were synthesized as described for 1 and 2 above and separated by preparative TLC (85:1 CHCl3/ethyl acetate, Rf = 0.81 (3) and 0.85 (4)) with 25% yield. 8-Aminoquinoline was obtained from Oakwood Chemicals, Estill, SC, USA.

3.4.1. Compound 3

Molecules 26 01008 i019
1H-NMR (500 MHz, CDCl3) δ 10.06 (s, 1H, NH), 8.78 (m, 1H, H-2), 8.76 (d, J = 7.8 Hz, 1H, H-4), 8.14 (dd, J = 8.2, 1.7 Hz, 1H, H-5), 7.53 (t, J = 7.8 Hz, 1H, H-6), 7.48 (dd, J = 8.3, 1.6 Hz, 1H, H-7), 7.43 (dd, J = 8.3, 4.2 Hz, 1H, H-3), 5.67 (s, 1H, H-3′), 3.34 (m, 1H, H-10′), 3.16 (dd, J = 15.3, 1.8 Hz, 1H, H-10′), 2.45 (m, 1H, H-4′), 2.40 (m, 1H, H-6′), 2.32 (m, 1H, H-4′), 2.22 (d, J = 5.6 Hz, 1H, H-1′), 2.12 (s, 1H, H-5′), 1.48 (d, J = 8.8 Hz, 1H, H-6′), 1.26 (s, 3H, H-9′), 0.85 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 169.6 (C=O), 148.3 (C-2), 142.2 (C-2′), 138.6 (C-8), 136.4 (C-5), 134.7 (C-8a), 128.0 (C-4a), 127.7 (C-6), 122.6 (C-7), 121.7 (C-3), 121.5 (C-4), 116.5 (C-3′), 47.2 (C-1′), 46.0 (C-10′), 40.5 (C-5′), 38.3 (C-6′), 32.1 (C-4′), 31.8 (C-7′), 26.2 (C-9′), 21.1 (C-8′). HRMS: [M + H]+: 307.1810 m/z calculated for C20H23N2O, found 307.1804 m/z.

3.4.2. Compound 4

Molecules 26 01008 i020
1H-NMR (500 MHz, CDCl3) δ 9.74 (s, 1H, NH), 8.85 (d, J = 7.8 Hz, 1H, H-2), 8.80 (dd, J = 4.3, 1.7 Hz, 1H, H-4), 8.16 (dd, J = 8.3, 1.7 Hz, 1H, H-5), 7.54 (t, J = 7.9 Hz, 1H, H-6), 7.49-7.47 (m, 1H, H-7), 7.45 (dd, J = 8.2, 4.2 Hz, 1H, H-3), 5.80 (s, 1H, H-3′), 3.55–3.48 (m, 1H, H-10′), 2.92–2.82 (m, 1H, H-10′), 2.57 (m, 1H, H-6′), 2.42 (m, 1H, H-4′), 2.09 (m, 1H, H-1′), 1.99 (m, 1H, H-5′), 1.93 (m, 1H, H-4′), 1.46 (d, J = 9.9 Hz, 1H, H-6′), 1.29 (s, 3H, H-9′), 0.81 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 166.4 (C=O), 165.5 (C-2), 148.1 (C-2′), 138.6 (C-8), 136.5 (C-5), 135.3 (C-8a), 128.1 (C-4a), 127.7 (C-6), 121.6 (C-7), 121.1 (C-3), 116.1 (C-3′), 115.9 (C-4), 54.0 (C-1′), 41.2 (C-5′), 40.7 (C-10′), 27.4 (C-6′), 26.3 (C-9′), 23.9 (C-7′), 22.5 (C-4′), 22.3 (C-8′). HRMS: [M + H]+: 307.1810 m/z calculated for C20H23N2O, found 307.1803 m/z.

3.5. Synthesis of 2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl)-N-(quinolin-4-yl) Acetamide (5)

Compound 5 was synthesized as described for 1 and 2 above and separated by preparative TLC (20:1 CHCl3/ethyl acetate, Rf = 0.55) and 25% yield. 4-Aminoquinoline (95% purity) was obtained from Ark Pharm, Inc., Arlington Heights, IL, USA.
Molecules 26 01008 i021
1H-NMR (500 MHz, CDCl3) δ 8.85 (d, J = 5.2 Hz, 1H, H-2), 8.34 (d, J = 5.2 Hz, 1H, H-3), 8.14 (d, J = 8.4 Hz, 1H, H-5), 7.74 (m, 1H, H-6), 7.73-7.69 (m, 1H, H-8), 7.58 (t, J = 7.58 Hz, 1H, H-7), 5.77 (s, 1H, H-3′), 3.37 (d, J = 15.5 Hz, 1H, H-10′), 3.18 (d, J = 15.4 Hz, 1H, H-10′), 2.53–2.48 (m, 1H, H-6′), 2.47 (m, 1H, H-5′), 2.42–2.36 (m, 1H, H-4′), 2.25–2.16 (m, 2H, H-4′), 2.19 (s, 1H, H-6′), 1.29 (s, 3H, H-9′), 1.26 (m, 1H, H-6′), 0.86 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 169.4 (C=O), 151.2 (C-2), 142.7 (C-2′), 140.3 (C-4a), 130.5 (C-8a), 129.4 (C-5), 126.6 (C-6), 123.6 (C-7), 119.7 (C-3), 118.6 (C-3′), 113.3 (C-8), 110.3 (C-4), 47.0 (C-1′), 45.8 (C-10′), 40.3 (C-5′), 38.2 (C-6′), 32.5 (C-4′), 31.7 (C-7′), 26.1 (C-9′), 21.1 (C-8′). HRMS: [M − H]-: 305.1654 m/z calculated for C20H21N2O, found 305.1655 m/z.

3.6. Synthesis of N-(7-chloroquinolin-4-yl)-2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl) Acetamide (6)

Compound 6 was synthesized as described for 1 and 2 above and separated by preparative TLC (20:1 hexane/ethyl acetate, Rf = 0.78) and 36% yield. 7-Chloro-4-aminoquinoline (95% purity) was obtained from Life Chemicals, Ontario, Canada.
Molecules 26 01008 i022
1H-NMR (500 MHz, CDCl3) δ 8.83 (d, J = 5.1 Hz, 1H, H-2), 8.31 (d, J = 5.2 Hz, 1H, H-3), 8.11 (s, J = 2.1 Hz, 1H, H-8), 7.65 (d, J = 9.0 Hz, 1H, H-5), 7.52 (dd, J = 9.0, 2.1 Hz, 1H, H-6), 5.79–5.75 (m, 1H, H-3′), 3.37 (d, J = 16.6 Hz, 1H, H-10′), 3.18 (d, J = 17.5 Hz, 1H, H-10′), 2.54–2.49 (m, 1H, H-6′), 2.49–2.44 (m, 1H, H-5′), 2.38 (m, 1H, H-1′), 2.21 (m, 2H, H-4′), 1.29 (s, 3H, H-9′), 1.23 (d, J = 8.8 Hz, 1H, H-6′), 0.86 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 169.5 (C=O), 152.4 (C-2), 142.8 (C-2′), 138.7 (C-4a) 136.7 (C-7), 135.6 (C-8), 129.5 (C-8a), 127.6 (C-5), 123.9 (C-6), 120.5 (C-3), 118.3 (C-3′), 110.7 (C-4), 47.1 (C-1′), 45.9 (C-10′), 40.4 (C-5′), 38.3 (C-6′), 32.7 (C-4′), 31.9 (C-7′), 26.1 (C-9′), 21.2 (C-8′). HRMS: [M − H]: 339.1264 m/z calculated for C20H20ClN2O, found 339.1263 m/z.

3.7. Synthesis of quinolin-4-yl-2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl) Acetate (7)

Molecules 26 01008 i023
N-(3-Dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride (EDCI, Sigma-Aldrich, 99% purity, 171.7 mg, 0.90 mmol, 1.3 eq.), DMAP (4-Dimethylaminopyridine, 99%, ACROS Organics™, 98.5% purity, 25.7 mg, 0.21 mmol, 0.5 eq.), and DIEA (2.04 mmol, 0.36 mL, and 3eq.) were added to a solution of 4-hydroxyquinoline (100 mg, 0.68 mmol, 1.0 eq., Ark Pharm, Inc., 97% purity) and nopoic acid (0.42 mmol, 1.2 eq.) in tetrahydrofuran (THF, Fisher Scientific, 99.9%, 2.0 mL) at 0 °C and stirred for 15 min. [25]. The temperature was raised to 23 °C, and the reaction mixture was stirred for 3 h while monitoring by analytical TLC. The reaction appeared cloudy at first but after stirring for 1 h at room temperature, it became clear, with the formation of colorless solid precipitate. The reaction mixture was quenched with brine, extracted with EtOAc, concentrated, and separated by preparative TLC (20:1 CHCl3/ethyl acetate, Rf = 0.55) with 25% yield.
1H-NMR (500 MHz, CD3OD) δ 8.84 (d, J = 5.0 Hz, 1H, H-2), 8.05 (d, J = 8.5 Hz, 1H, H-8), 8.02 (d, J = 8.5, 1.4 Hz, 1H, H-5), 7.81 (ddd, J = 8.4, 6.9, 1.4 Hz, 1H, H-6), 7.64 (ddd, J = 8.2, 6.8, 1.2 Hz, 1H, H-7), 7.44 (d, J = 5.0 Hz, 1H, H-3), 5.97 (m, 1H, H-3′), 3.34–3.28 (m, 1H, H-10′), 2.82-2.70 (m, 2H, H-10′, H-1′), 2.52 (m, 1H, H-6′), 2.11 (m, 1H, H-5′), 2.02 (m, 1H, H-4′), 1.96 (m, 1H, H-4′), 1.48 (d, J = 10.2 Hz, 1H, H-6′), 1.34 (s, 3H, H-9′), 0.84 (s, 3H, H-8′). 13C-NMR (126 MHz, CD3OD) δ 177.2 (C=O), 164.3 (C-4a), 156.5 (C-2), 151.9 (C-7), 150.5 (C-2), 131.7 (C-8), 129.2 (C-8a), 124.2 (C-6), 122.8 (C-3), 114.6 (C-3′), 111.4 (C-4′), 42.1 (C-1′), 41.6 (C-10′), 34.7 (C-5′), 28.0 (C-6′), 26.7 (C-4′), 26.4 (C-7′), 26.1 (C-9′), 24.5 (C-8′). [M + H]+: 308.1650 m/z calculated for C20H22NO2, found 308.1651 m/z.

3.8. Synthesis of 7-chloroquinolin-4-yl-2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl) Acetate (8)

Compound 8 was synthesized as described for 7 above and separated by preparative TLC (20:1 hexane/ethyl acetate and Rf = 0.85) with 40% yield. 7-Chloro-4-hydroxyquinoline (99% purity) was obtained from Sigma-Aldrich.
Molecules 26 01008 i024
1H-NMR (500 MHz, (CD3)2CO) δ 8.95 (d, J = 4.8 Hz, 1H, H-2), 8.10 (d, J = 2.2 Hz, 1H, H-8), 8.06 (d, J = 9.0 Hz, 1H, H-5), 7.62 (dd, J = 8.9, 2.0 Hz, 1H, H-6), 7.49 (d, J = 4.9 Hz, 1H, H-3), 5.99 (m, 1H, H-3′), 3.31 (dd, J = 20.4, 9.2 Hz, 1H, H-10′), 2.84–2.78 (m, 1H, H-10′), 2.75 (m, 1H, H-1′), 2.51 (dt, J = 10.8, 5.8 Hz, 1H, H-6′), 2.14–2.10 (m, 1H, H-5′), 2.02 (ddt, J = 10.1, 3.9, 1.9 Hz, 1H, H-4′), 1.94 (m, 1H, H-4′), 1.49 (d, J = 10.1 Hz, 1H, H-6′), 1.33 (s, 3H, H-9′), 0.84 (s, 3H, H-8′). 13C-NMR (126 MHz, (CD3)2CO) δ 176.1 (C=O), 163.4 (C-4a), 155.3 (C-2), 153.4 (C-7), 151.2 (C-2′), 136.1 (C-8), 129.0 (C-8a), 128.3 (C-5), 124.5 (C-6), 122.2 (C-3), 114.6 (C-3′), 111.3 (C-4), 54.7 (C-1′), 41.7 (C-10′), 41.1(C-5′), 27.6 (C-6′), 26.2 (C-9′), 24.1 (C-4′), 23.6 (C-7′), 22.4 (C-8′). [M + H]+: 342.1261 m/z for calculated C20H21ClNO2, found 342.1259 m/z.

3.9. Synthesis of 7-chloro-2,8-dimethylquinolin-4-yl-2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl) Acetate (9)

Compound 9 was synthesized as described for 7 above and separated by preparative TLC (4:1 hexane/ethyl acetate, Rf = 0.89) with 39% yield. 7-Chloro-2,8-dimethyl-4-hydroxyquinoline (98% purity) was obtained from Sigma-Aldrich.
Molecules 26 01008 i025
1H-NMR (500 MHz, (CD3)2CO) δ 7.78 (d, J = 9.0 Hz, 1H, H-6), 7.50 (d, J = 8.9 Hz, 1H, H-5), 7.31 (s, 1H, H-3), 5.96 (m, 1H, H-3′), 3.34–3.27 (m, 1H, H-10′), 2.82 (s, 3H, C-3Me), 2.77 (m, 1H, H-10′) 2.74 (m, 1H, H-1′), 2.71 (s, 3H, C-8Me), 2.51 (m, 1H, H-6′), 2.11 (m, 1H, H-5′), 2.03–1.99 (m, 1H, H-4′), 1.97–1.91 (m, 1H, H-4′), 1.48 (d, J = 10.1 Hz, 1H, H-6′), 1.33 (s, 3H, H-9′), 0.83 (s, 3H, H-8′). 13C-NMR (126 MHz (CD3)2CO) δ 175.6 (C=O), 163.5 (C-4a), 160.8 (C-2), 155.6 (C-7), 149.6 (C-2′), 135.6 (C-8), 134.9 (C-8a), 127.6 (C-5), 120.9 (C-6), 120.5 (C-3), 114.6 (C-3′), 111.4 (C-4), 54.7 (C-1′), 41.7 (C-10′), 41.1 (C-5′), 27.6 (C-6′), 26.2 (C-9′), 25.8 (C-4′), 24.1 (C-2Me), 23.5 (C-7′), 22.4 (C-8′), 14.7 (C-8Me). [M + H]+: 370.1574 m/z for calculated C22H25ClNO2, found 370.1574 m/z.

3.10. Synthesis of 6-fluoro-2-(trifluoromethyl)quinolin-4-yl-2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2yl) Acetate (10)

Compound 10 was synthesized as described for 7 above and separated by preparative TLC (4.5:1 hexane/ethyl acetate, Rf = 0.78) with 33% yield. 6-Fluoro-4-hydroxy-2-(trifluoromethyl)quinoline (98% purity) was obtained from Sigma-Aldrich.
Molecules 26 01008 i026
1H-NMR (500 MHz, CDCl3) δ 8.25 (m, 1H, H-7), 7.78 (s, 1H, H-3), 7.67 (m, 1H, H-8), 7.60 (s, 1H, H-5), 5.91 (s, 1H, H-3′), 3.33 (dd, J = 20.5, 9.3 Hz, 1H, H-10′), 2.88–2.73 (m, 1H, H-10′), 2.71 (t, J = 5.4 Hz, 1H, H-1′), 2.50 (m, 1H, H-6′), 2.16 (m, 1H, H-5′), 2.05–1.93 (m, 2H, H-4′), 1.47 (d, J = 10.2 Hz, 1H, H-6′), 1.34 (s, 3H, H-9′), 0.85 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 177.7 (C=O), 162.4 (C-6), 146.2 (C-2), 142.2 (C-2′), 133.1 (C-8), 133.0 (C-4a), 124.2 (CF3), 121.9 (C-7), 121.7 (C-8a), 110.1 (C-3′), 109.8 (C-5), 105.8 (C-3), 105.6 (C-4), 54.4 (C-1′), 41.4 (C-10′), 40.4 (C-5′), 27.4 (C-6′), 26.2 (C-9′), 23.7 (C-4′), 23.4 (C-7′), 22.4 (C-8′). [M + Na]+: 416.1250 m/z calculated for C21H19F4NO2Na, found: 416.1245 m/z.

3.11. Synthesis of 7-chloro-8-methylquinolin-4-yl-2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl) Acetate (11)

Compound 11 was synthesized as described for 7 above and separated by preparative TLC (5:1 hexane/ethyl acetate, Rf = 0.82) with 38% yield. 7-Chloro-4-hydroxy-8-methylquinoline (98% purity) was obtained from Sigma-Aldrich.
Molecules 26 01008 i027
1H-NMR (500 MHz, CDCl3) δ 8.92 (d, J = 4.9 Hz, 1H, H-2), 7.79 (d, J = 8.9 Hz, 1H, H-6), 7.51 (d, J = 9.0 Hz, 1H, H-5), 7.33 (d, J = 4.9 Hz, 1H, H-3), 5.90 (t, J = 2.3 Hz, 1H, H-3′), 3.36–3.27 (m, 1H, H-10′), 2.88 (s, 3H, C-8Me), 2.76 (m, 1H, H-10′), 2.68 (t, J = 5.3 Hz, 1H, H-1′), 2.48 (m, 1H, H-6′), 2.13 (s, 1H, H-5′), 2.03-1.98 (m, 1H, H-4′), 1.93 (m, 1H, H-4′), 1.46 (d, J = 10.1 Hz, 1H, H-6′), 1.25 (s, 3H, H-9′), 0.84 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 175.8 (C=O), 163.0 (C-4a), 154.6 (C-2), 150.5 (C-7), 149.8 (C-2′), 135.4 (C-8), 135.0 (C-8a), 128.0 (C-5), 121.0 (C-6), 119.9 (C-3), 112.8 (C-3′), 110.5 (C-4), 54.2 (C-1′), 41.1 (C-10′), 40.35 (C-5′), 31.92 (C-7′), 27.24 (C-6′), 26.02 (C-9′), 23.53 (C-4′), 22.23 (C-8′), 14.77 (C-8Me). [M + Na]+: 378.1237 m/z for calculated for C21H22ClNO2Na found 378.1234 m/z.

3.12. Synthesis of 2,8-bis(trifluoromethyl)quinolin-4-yl-2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl) Acetate (12)

Compound 12 was synthesized as described for 7 above and separated by preparative TLC (4.5:1 hexane/ethyl acetate, Rf = 0.76) with 31% yield. 2,8-Bis(trifluoromethyl)-4-quinolinol (99% purity, Acros Organics) was obtained from Thermo Fisher Scientific, Waltham, MA.
Molecules 26 01008 i028
1H-NMR (500 MHz, CDCl3) δ 8.28 (d, J = 8.6 Hz, 1H, H-7), 8.18 (d, J = 7.4 Hz, H-5), 7.84 (s, 1H, H-3), 7.74–7.67 (m, 1H, H-6), 5.92 (s, 1H, H-3′), 3.38–3.28 (m, 1H, H-10′), 2.80 (m, 1H, H-10′), 2.71 (m, 1H, H-1′), 2.50 (m, 1H, H-6′), 2.16 (m, 1H, H-5′), 2.08–2.00 (m, 1H, H-4′), 1.96 (m, 1H, H-4′), 1.47 (d, J = 10.3 Hz, 1H, H-6′), 1.34 (s, 3H, H-9′), 0.85 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 178.0 (C=O), 162.4 (C-6), 156.2 (C-2), 149.3 (C-8a), 145.4 (C-2′), 129.6 (C-8), 129.6 (C-4a), 127.21 (C-4), 126.3 (C-5), 124.7 (CF3), 124.0 (CF3), 122.6 (C-7), 122.5 (C-3), 110.1 (C-3′), 54.5 (C-1′), 41.4 (C-10′), 40.4 (C-5′), 27.4 (C-6′), 26.2 (C-9′), 23.7 (C-4′), 23.5 (C-8′), 22.4 (C-8′). [M + Na]+: 466.1218 m/z calculated for C22H19F6NO2Na, found 466.1220 m/z.

3.13. Synthesis of 2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl)ethyl-7-chloro-2-methylquinoline-4-carboxylate (13)

EDCI (112.1 mg, 0.58 mmol, and 1.3 eq.) was added to a solution of 7-chloro-2 methyl-4-quinolinecarboxylic acid (99.7 mg, 0.45 mmol, 1.0 eq., Sigma-Aldrich, 98% purity) in THF (1.0 mL) at 0 °C and stirred for 15 min. Nopol (90.0 mg, 0.54 mmol, 1.2 eq.) in THF (1.0 mL) was added, followed by DIEA (2.04 mmol, 0.36 mL, 3 eq.). The reaction mixture was stirred at room temperature for 3 h while monitoring by analytical TLC. The reaction mixture was quenched with brine, extracted with EtOAc, concentrated, and separated by preparative TLC (3:1 hexane/ethyl acetate, Rf = 0.92) with 35% yield.
Molecules 26 01008 i029
1H-NMR (500 MHz, CDCl3) δ 8.69 (dd, J = 9.1, 1.5 Hz, 1H, H-6), 8.06 (d, J = 1.9 Hz 1H, H-8), 7.78 (d, J = 1.6 Hz 1H, H-3), 7.51 (d, J = 9.1 Hz, 1H, H-5), 5.40 (m, 1H, H-3′), 4.45 (m, 2H, H-11′), 2.77 (d, J = 1.7 Hz, 3H, C-2Me), 2.49 (s, 2H, H-10′), 2.43–2.38 (m, 1H, H-6′), 2.33-2.20 (m, 2H, H-4′), 2.15–2.09 (m, 2H, H-5′, H-1′), 1.28 (s, 3H, H-9′), 1.18 (d, J = 8.6 Hz, 1H, H-6′), 0.85 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 166.1 (C=O), 160.0 (C-3), 149.5 (C-2), 144.0 (C-2′), 135.8 (C-8), 135.3 (C-8a), 128.3 (C-5), 128.2 (C-6), 127.1 (C-4), 123.6 (C-4a), 122.0 (C-3), 119.5 (C-3′), 64.3 (C-11′), 45.9 (C-10′), 40.8 (C-1′), 38.2 (C-5′), 36.1 (C-7′), 31.9 (C-6′), 31.6 (C-4′), 26.4 (C-9′), 25.3 (C-2Me), 21.3 (C-8′). [M + H]+: 370.1574 m/z calculated for C22H25ClNO2, found 370.1575 m/z.

3.14. Synthesis of 2-((1S,5R)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl)ethyl-2-chloroquinoline-3-carboxylate (14)

Compound 14 was synthesized as described for 13 above and separated by preparative TLC preparative TLC by preparative TLC (40:1 CHCl3/Et2O, Rf = 0.72) with 50% yield. The acid, 2-chloroquinoline-3-carboxylic acid (97% purity) was obtained from Sigma-Aldrich.
Molecules 26 01008 i030
1H-NMR (500 MHz, CD3OD) δ 8.76 (s, 1H, H-4), 8.01 (d, J = 8.1 Hz, 1H, H-5), 7.94 (d, J = 8.5 Hz, 1H, H-8), 7.88 (dd, J = 8.5, 6.9 Hz, 1H, H-7), 7.68 (t, J = 7.6 Hz, 1H, H-6), 5.39 (s, 1H, H-3′), 4.38 (qt, J = 10.9, 6.8 Hz, 2H, H-11′), 2.46 (m, 2H, H-10′), 2.40 (dt, J = 8.7, 5.7 Hz, 1H, H-6′), 2.30-2.17 (m, 2H, H-4′), 2.17-2.12 (m, 1H, H-5′), 2.06 (m, 1H, H-1′), 1.27 (s, 3H, H-9′), 1.17 (d, J = 8.6 Hz, 1H, H-6′), 0.84 (s, 3H, H-8′). 13C-NMR (126 MHz, CD3OD) 165.70 (C=O), 149.2 (C-2), 148.4 (C-4a), 145.5 (C-3), 143.0 (C-3′), 134.1 (C-6), 129.9 (C-8), 129.3 (C-8a), 128.7 (C-4), 127.3 (C-5), 125.9 (C-7), 120.2 (C-3′), 65.3 (C-11′), 46.9 (C-10′), 42.0 (C-5′), 38.9 (C-1′), 36.9 (C-6′), 32.6 (C-7′), 32.3 (C-4′), 26.7 (C-9′), 21.6 (C-8′). [M + H]+: 356.1417 m/z calculated for C21H23ClNO2, found 356.1409 m/z.

3.15. Synthesis of 2-((1R,5S)-6,6-dimethylbicyclo[3.1.1]hept-2-en-2-yl)ethyl-1H-indole-2-carboxylate (15)

Compound 15 was synthesized as described for 13 above and separated by preparative TLC (40:1 CHCl3/Et2O, Rf = 0.72) with 50% yield. Indole 2-carboxylic acid (98% purity) was obtained from Sigma-Aldrich.
Molecules 26 01008 i031
1H-NMR (500 MHz, CD3OD) δ 7.61 (d, J = 8.1 Hz, 1H, H-7), 7.43 (d, J = 8.4 Hz, 1H, H-4), 7.24 (t, J = 7.7 Hz, 1H, H-5), 7.11 (s, 1H, H-3), 7.06 (t, J = 7.6 Hz, 1H, H-6), 5.39 (s, 1H, H-3′), 4.38–4.27 (m, 2H, H-11′), 2.47–2.39 (m, 3H, H-10′, H-6′), 2.31–2.19 (m, 2H, H-4′), 2.17 (t, J = 5.6 Hz, 1H, H-5′), 2.07 (s, 1H, H-1′), 1.28 (s, 3H, H-9′), 1.19 (d, J = 8.6 Hz, 1H, H-6′), 0.86 (s, 3H, H-8′). 13C-NMR (126 MHz, CD3OD) δ 163.5 (C=O), 145.8 (C-3′), 139.0 (C-7a), 128.6 (C-3a), 126.0 (C-2), 123.1 (C-4), 121.3 (C-6), 120.2 (C-5), 113.3 (C-2′), 109.2 (C-7), 64.2 (C-11′), 47.0 (C-10′), 42.0 (C-1′), 40.0 (C-5′), 37.1 (C-7′), 32.6 (C-6′), 32.4 (C-4′), 26.7 (C-9′), 24.2 (C-9), 21.6 (C-8) HRMS: [M + H]+: 310.1807 m/z calculated for C20H24NO2, found 310.1806 m/z.

3.16. Synthesis of 2-((1S,2S,4R,6S)-7,7-dimethyl-3-oxatricyclo[4.1.1.0 2,4]octan-2-yl)ethyl-7-chloro-2-methylquinoline-4-carboxylate (16)

Nopol (500 mg, 3.0 mmol, 1.eq.) was added to a round 100 mL bottom flask, followed by 4.0 mL of ethyl acetate, 4.0 mL of deionized H2O, and 8.0 mL of acetone, and stirred at room temperature. NaHCO3(s) (9.0 mmol, 756 mg, 3.0 eq.) was added to the mixture and stirred for 15 min. Oxone salt (3.6 mmol, 2.4 g, 1.2 eq.), dissolved in 1.0 mL of water, was added dropwise for 1 h, and analytical TLC was used to monitor the reaction for 2.5 h [26]. The reaction was quenched with 3.0 mL of H2O. The organic layer was extracted with 15.0 mL CHCl3 (3 times), dried over MgSO4(s), and concentrated. The crude colorless solid obtained was separated by preparative TLC (10:1 CHCl3/ethyl acetate, Rf = 0.37) with 40% yield of nopol oxide. EDCI (112.1 mg, 0.58 mmol, 1.3 eq.), DMAP (27.5 mg, 0.23 mmol, 0.5 eq.), and DIEA (2.04 mmol, 0.36 mL, 3 eq.) were added to a solution of 7-chloro-2-methyl-4-quinolinecarboxylic acid (99.7 mg, 0.45 mmol, 1.0 eq., Sigma-Aldrich, 98%) and nopol oxide (151.9 mg, 0.54 mmol, and 1.2 eq.) in THF (2.0 mL) at 0 °C and stirred for 15 min. The temperature was raised to 23 °C, and the mixture was stirred for 3 h while monitoring by analytical TLC, quenched with brine, extracted with EtOAc, and separated by preparative TLC (4.5:1 hexane/ethyl acetate, Rf = 0.59) with 49% yield.
Molecules 26 01008 i032
1H-NMR (500 MHz, CDCl3) δ 8.68 (d, J = 9.1 Hz, 1H, H-5), 8.06 (d, J = 2.2 Hz, 1H, H-8), 7.79 (s, 1H, H-3), 7.53 (d, J = 9.1, 2.2 Hz, 1H, H-6), 4.57–4.44 (m, 2H, H-11′), 3.21 (d, J = 4.2, 1.5 Hz, 1H, H-3′), 2.78 (s, 3H, C-2Me), 2.29 (dt, J = 14.8, 6.0 Hz, 1H, H-6′), 2.16 (s, 1H, H-5′), 2.08–2.02 (m, 2H, H-4′, H-1′), 1.94 (m, 1H, H-10′), 1.77 (m, 1H, H-10′), 1.67 (d, J = 9.8 Hz, 1H, H-6′), 1.32 (s, 3H, H-9′), 0.98 (s, 3H, H-8′). 13C-NMR (126 MHz, CDCl3) δ 166.0 (C=O), 160.0 (C-3), 149.5 (C-2), 135.9 (C-8), 135.1 (C-8a), 128.3 (C-5), 128.2 (C-6), 127.1 (C-4), 123.6 (C-4a), 121.9 (C-3), 61.9 (C-3′), 62.0 (C-11′), 55.5 (C-2′), 43.7 (C-10′), 40.8 (C-1′), 40.0 (C-5′), 34.0 (C-7′), 27.7 (C-6′), 26.9 (C-4′), 25.9 (C-9′), 25.5 (C-2Me), 20.3 (C-8′). [M + H]+: 386.1523 m/z for calculated C22H25ClNO3, found 386.1525 m/z.

3.17. Plasmodium falciparum Assay

a. Pf3D7: 100 μL of complete media (Sigma-Aldrich’s RPMI 1640, 25 mM HEPES, 10 μg/mL gentamicin, 0.5 mM hypoxanthine supplemented with NaHCO3 and 0.5% Albumax II, 37 °C) containing compounds (50–0.3 µM) or 100 µM chloroquine were added to 96 well plates in triplicates. This was followed by the addition of 100 µL erythrocytic asexual culture of Pf3D7 (10% hematocrit and 0.25% ring stage parasitemia), maintained at atmospheric conditions of 1% oxygen, 5% carbon dioxide, and 94% nitrogen, and allowed to grow for 96 h at 37 °C. The cultures were then frozen at −80 °C overnight and thawed at 37 °C for 4 h. Aliquots (100 µL) from each well were transferred to black 96-well plates, 100 μL of 2× SYBR Green (Molecular Probes) in lysis buffer (20 mM Tris at pH 7.5, 5 mM EDTA, 0.008% saponin, 0.08% Triton X-100) was added to each well and mixed thoroughly with a pipette. The plates were incubated in the dark at room temperature for 1 h and read at λex 485 nm and λem 530 nm [27]. Infected and untreated red blood cells (iRBC) cultures served as negative control. Percent inhibitory activities were calculated using (=100 − 100*((Sample − 100 µM CQ)/(iRBC − 100 µM CQ)) and used to plot sigmoidal inhibition curves to obtain the EC50 values.
b. PfNH54 and PfK1: The growth inhibitory activities of the compounds against erythrocytic stages of PfNH54 and PfK1 were determined using the [3H]-hypoxanthine incorporation assay as previously reported [28]. Parasites in RPMI 1640 medium containing 5% Albumax (without hypoxanthine) were treated with compounds (50–0.3 µM) in 96 well plates for 48 h at 37 °C in 92% N2, 5% CO2, and 3% O2. 3H-hypoxanthine (0.5 μCi) was added to each well, incubated for an additional 24 h, harvested onto glass-fiber filters, and washed with distilled water. Radioactivity at each compound concentration was counted using a Betaplate™ liquid scintillation counter (Wallac, Zurich), and the results were recorded as counts per minute (CPM) per well. The CPMs were expressed as a percentage of the CPMs from the untreated controls and used to estimate the EC50 from sigmoidal inhibition curves. Artemisinin and CQ were used as positive controls.

3.18. Cytotoxicity Assay

Human hepatocarcinoma cell line (Hep G2, ATCC CRL-11997TM) was used for cytotoxicity studies as previously described [29]. The cells were grown in complete medium (DMEM:F12 containing l-glutamine and sodium bicarbonate, 10% FBS, 1% penicillin/streptomycin, Thermo Fisher Scientific) at 37 °C in 5% CO2 environment. Cells (198 µL, 5 × 105/mL) were seeded into 96-well plates and incubated overnight. The cells were treated with the compounds (160–1.25 µM, in triplicates) or DMSO (1%) for 72 h. The cell medium was removed and replaced with DMEM:F12 medium containing MTT (0.5 mg/mL) was added to the cells and incubated for 1.5 h. The MTT-containing medium was gently removed and replaced with DMSO (200 µL/well). The contents of each well were then repeatedly mixed with a multichannel pipette and incubated for 10 min. The plates were read at 570 nm. DMSO-treated (1%) cells were used as negative assay control and SDS (10%) was used as assay positive control. Percent cell viabilities were expressed as a percentage of the mean viability of DMSO-treated (1%) cells.

Supplementary Materials

The following are available online. 1H, 13C, and HRMS Data (Figure S1.1a–S1.16b).

Author Contributions

Conceptualization, R.J.N. and I.V.O.; methodology, R.J.N., H.Z., J.T.C., O.C, I.V.O.; formal analysis, R.J.N., H.Z., J.T.C., O.C., I.V.O.; resources, I.V.O.; investigation, R.J.N., H.Z., J.T.C., O.C., I.V.O.; data curation, R.J.N., H.Z., J.T.C., O.C., I.V.O.; writing—original draft preparation, R.J.N. and I.V.O.; writing—review and editing, R.J.N. and I.V.O.; supervision, I.V.O.; project administration, I.V.O.; funding acquisition, I.V.O. All authors have read and agreed to the published version of the manuscript.”

Funding

This research was partly funded by the US National Institutes of Health, SC3GM122629 and G12MD007581. J.T.C. was supported by The National Science Foundation (NSF) HBCU-RISE Program (HRD-1547836). O.C. was supported in part by the American Association of University Women (AAUW).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or supplementary material.

Acknowledgments

We thank Marcel Kaiser (Swiss Tropical and Public Health Institute, Basel, Switzerland) for technical assistance with the Plasmodium assays.

Conflicts of Interest

The authors declare no conflict of interests.

Sample Availability

Samples of the compounds 116 are available from the authors.

References

  1. Centers for Disease Control and Prevention (CDC). Malaria. Available online: https://www.cdc.gov/parasites/malaria/index.html (accessed on 12 January 2020).
  2. World Health Organization (WHO). Global Malaria Programme. Available online: https://www.who.int/teams/global-malaria-programme (accessed on 12 January 2020).
  3. Witkowski, A.B.; Amaratunga, C.; Beghain, J.; Langlois, A.C.; Khim, N.; Kim, S.; Duru, V.; Bouchier, C.; Ma, L.; Lim, P.; et al. A Molecular Marker of Artemisinin-resistant Plasmodium falciparum malaria. Nature 2014, 505, 50–55. [Google Scholar] [CrossRef]
  4. Slater, A.F.; Swiggard, W.J.; Orton, B.R.; Flitter, W.D.; Goldberg, D.E.; Cerami, A.; Henderson, G.B. An iron-carboxylate bond links the heme units of malaria pigment. Proc. Natl. Acad. Sci. USA 1991, 88, 325–329. [Google Scholar] [CrossRef] [Green Version]
  5. Sullivan, D.J.; Matile, H.; Ridley, R.; Goldberg, D.E. A Common Mechanism for Blockade of Heme Polymerization by Antimalarial Quinolines. J. Biol. Chem. 1998, 47, 31103–31107. [Google Scholar] [CrossRef] [Green Version]
  6. Sullivan, D.J., Jr.; Gluzman, I.Y.; Goldberg, D.E. Plasmodium hemozoin formation mediated by histidine-rich proteins. Science 1996, 271, 219–222. [Google Scholar] [CrossRef] [PubMed]
  7. Martiney, J.A.; Cerami, A.; Slater, A.F. Inhibition of Hemozoin formation in Plasmodium falciparum trophozoite extracts by heme analogs: Possible Implication in the Resistance to Malaria conferred by the β-thalassemia trait. Mol. Med. 1996, 2, 236–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Egan, T.J.; Combrinck, J.M.; Egan, J.; Hearne, G.R.; Marques, H.M.; Ntenteni, S.; Sewell, B.T.; Smith, P.J.; Taylor, D.; Van Schalkwyk, D.A.; et al. Fate of Haem Iron in the malaria parasite Plasmodium falciparum. Biochem. J. 2002, 365, 343–347. [Google Scholar] [CrossRef] [PubMed]
  9. Wicht, K.J.; Mok, S.; Fidock, D.A. Molecular mechanisms of drug resistance in Plasmodium falciparum malaria. Annu. Rev. Microbiol. 2020, 74, 431–454. [Google Scholar] [CrossRef] [PubMed]
  10. Rawe, S.L.; McDonnell, C. The cinchona alkaloids and the aminoquinolines. In Antimalarial Agents; Patrick, G.L., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 65–98. [Google Scholar]
  11. Kim, J.; Tan, Y.Z.; Wicht, K.J.; Erramilli, S.K.; Dhingra, S.K.; Okombo, J.; Vendome, J.P.; Hagenah, L.M.; Giacometti, S.I.; Potter, C.S.; et al. Structure and Drug Resistance of the Plasmodium falciparum Transporter PfCRT. Biophys. J. 2020, 118, 523a. [Google Scholar] [CrossRef]
  12. Sá, J.M.; Kaslow, S.R.; Barros, R.R.M.; Brazeau, N.F.; Parobek, C.M.; Tao, D.; Salzman, R.E.; Gibson, T.J.; Velmurugan, S.; Krause, M.A.; et al. Plasmodium vivax chloroquine resistance links to pvcrt transcription in a genetic cross. Nat. Commun. 2019, 10, 4300. [Google Scholar] [CrossRef]
  13. Camarda, G.; Jirawatcharadech, P.; Priestley, R.S.; Saif, A.; March, S.; Wong, M.H.; Leung, S.; Miller, A.B.; Baker, D.A.; Alano, P.; et al. Antimalarial activity of primaquine operates via a two-step biochemical relay. Nat. Commun. 2019, 10, 3226. [Google Scholar] [CrossRef] [PubMed]
  14. Fasinu, P.S.; Nanayakkara, N.D.; Wang, Y.H.; Chaurasiya, N.D.; Herath, H.B.; McChesney, J.D.; Avula, B.; Khan, I.; Tekwani, B.L.; Walker, L.A. Formation primaquine-5,6-orthoquinone, the putative active and toxic metabolite of primaquine via direct oxidation in human erythrocytes. Malar. J. 2019, 18, 30. [Google Scholar] [CrossRef]
  15. Van Zyl, R.L.; Seatlholo, S.T.; van Vuuren, S.F. The biological activities of 20 nature identical essential oil constituents. J. Essent. Oil Res. 2006, 18, 129–133. [Google Scholar] [CrossRef]
  16. Bharate, S.B.; Singh, I.P. A two-step biomimetic synthesis of antimalarial robustadials A and B. Tetrahedron Lett. 2006, 47, 7021–7024. [Google Scholar] [CrossRef]
  17. Il’ina, L.; Volcho, K.; Korchagina, D.; Salakhutdinov, N.; Barkhash, V. Transformations of Epoxide Derived from Nopol over Askanite-Bentonite Clay. Russ. J. Org. Chem. 2019, 40, 1432–1436. [Google Scholar] [CrossRef]
  18. Cabrera, M.; Cui, L. In vitro activities of primaquine-schizonticide combinations on asexual blood stages and gametocytes of Plasmodium falciparum. Antimicrob. Agents Chemother. 2015, 59, 7650–7656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Boudhar, A.; Ng, X.W.; Loh, C.Y.; Chia, W.N.; Tan, Z.M.; Nosten, F.; Dymock, B.W.; Tan, K.S.W. Overcoming chloroquine resistance in malaria: Design, synthesis, and structure-activity relationships of novel hybrid compounds. Antimicrob. Agents Chemother. 2016, 60, 3076–3089. [Google Scholar] [CrossRef] [Green Version]
  20. Richards, S.N.; Nash, M.N.; Baker, E.S.; Webster, M.W.; Lehane, A.M.; Sharik, S.H.; Martin, R.E. Molecular Mechanisms for Drug Hypersensitivity Induced by the Malaria Parasite’s Chloroquine Resistance Transporter. PLoS Pathog. 2016, 12, e1005725. [Google Scholar] [CrossRef] [Green Version]
  21. Reddy, M.D.; Kobori, H.; Mori, T.; Wu, J.; Kawagishi, H.; Watkins, E.B. Gram-scale, stereoselective synthesis and biological evaluation of (+)-armillariol C. J. Nat. Prod. 2017, 80, 2561–2565. [Google Scholar] [CrossRef]
  22. Yang, D.; Jiao, G.S.; Yip, Y.C.; Wong, M.K. Diastereoselective epoxidation of cyclohexene derivatives by dioxiranes generated in situ. Importance of steric and field effects. J. Org. Chem. 1999, 64, 1635–1639. [Google Scholar] [CrossRef] [PubMed]
  23. Hashimoto, N.; Kanda, A. Practical and environmentally friendly epoxidation of olefins using oxone. Org. Process Res. Dev. 2002, 6, 405–406. [Google Scholar] [CrossRef]
  24. Caprino, L. 1-Hydroxy-7-azabenzotriazole. An efficient peptide coupling additive. J. Am. Chem. Soc. 1993, 115, 4397–4398. [Google Scholar] [CrossRef]
  25. Park, Y.; Fei, X.; Yuan, Y.; Lee, S.; Joonseong, H.; Jean, S.; Jung, J.; Seo, S. Chemoselective acylation of 2-amino-8-quinolinol in the generation of C2-amides or C8-esters. RSC Adv. 2017, 7, 41955–41961. [Google Scholar] [CrossRef] [Green Version]
  26. Denmark, S.; Forbes, D.C.; Hays, D.S.; DePue, J.S.; Wilde, R.G. Catalytic Epoxidation of Alkenes with Oxone. J. Org. Chem. 1997, 62, 1397–1407. [Google Scholar] [CrossRef] [PubMed]
  27. Smilkstein, M.; Sriwilaijaroen, N.; Kelly, J.X.; Wilairat, P.; Riscoe, M. Simple and inexpensive fluorescence-based technique for high-throughput antimalarial drug screening. Antimicrob. Agents Chemother. 2004, 48, 1803–1806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Matile, H.; Pink, J.R.L. Plasmodium falciparum malaria parasite cultures and their use in immunology. In Immunological Methods; Lefkovits, I., Pernis, B., Eds.; Academic Press: San Diego, CA, USA, 1990; pp. 221–234. [Google Scholar]
  29. Zhang, H.; Collins, J.; Nyamwihura, R.; Ware, S.; Kaiser, M.; Ogungbe, I. Discovery of a quinoline-based phenyl sulfone derivative as an antitrypanosomal agent. Bioorg. Med. Chem. Lett. 2018, 28, 1647–1651. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthesis of compounds 116. Reagent and conditions: (a) Jones reagent: Acetone, H2O, 50 °C, 12 h, 33–45% (b) DMF, HBTU, HOBT, DIEA, RNH2, r.t, 16 h, 20–40% (c) THF, EDCI, DMAP, ROH, DIEA, 0 °C–r.t, 3 h, 25–45% (d) THF, EDCI, DMAP, RCOOH, DIEA, 0 °C–r.t, 3 h, 45–55% (e) 2:1 ethyl acetate/acetone, NaHCO3, r.t, 5 min, then, Oxone® salt (KHSO5, K2SO4) in H2O, 1 h, 40%.
Scheme 1. Synthesis of compounds 116. Reagent and conditions: (a) Jones reagent: Acetone, H2O, 50 °C, 12 h, 33–45% (b) DMF, HBTU, HOBT, DIEA, RNH2, r.t, 16 h, 20–40% (c) THF, EDCI, DMAP, ROH, DIEA, 0 °C–r.t, 3 h, 25–45% (d) THF, EDCI, DMAP, RCOOH, DIEA, 0 °C–r.t, 3 h, 45–55% (e) 2:1 ethyl acetate/acetone, NaHCO3, r.t, 5 min, then, Oxone® salt (KHSO5, K2SO4) in H2O, 1 h, 40%.
Molecules 26 01008 sch001
Table 1. Antiplasmodial activity of compounds 116.
Table 1. Antiplasmodial activity of compounds 116.
CompoundRPf3D7
EC50 (µM) (SI) a
PfNF54
EC50 (µM) (SI) a
PfK1
EC50 (µM) (SI) a
Hep G2
CC50 (µM)
1 Molecules 26 01008 i0015.25 ± 0.70 (13.4)ndnd70.8 ± 1.26
2 Molecules 26 01008 i0022.72 ± 0.28 (>58.8)ndnd>160
3 Molecules 26 01008 i0032.19 ± 0.15 (>73.1)>50>50>160
4 Molecules 26 01008 i0042.57 ± 0.02 (>62.3)>50>50>160
5 Molecules 26 01008 i00526.4 ± 1.70 (0.2)23.4 ± 0.14 (0.2)13.9 ± 2.1 (0.3)3.98 ± 0.7
6 Molecules 26 01008 i00617.1 ± 1.80 (0.8)21.37 ± 1.70 (0.7)7.84 ± 0.57(1.8)14.26 ± 1.2
7 Molecules 26 01008 i007>50>50>28>160
8 Molecules 26 01008 i008>5023.27 ± 3.78 (>6.8)0.164 ± 0.06 (>976)>160
9 Molecules 26 01008 i0096.15 ± 0.98 (>26)ndnd>160
10 Molecules 26 01008 i01021.39 ± 5.06 (2.3)ndnd48.50 ± 1.12
a Selectivity index (CC50/EC50); nd = not determined.
Table 2. Antiplasmodial activity of compounds 116.
Table 2. Antiplasmodial activity of compounds 116.
CompoundRPf3D7
EC50 (µM) (SI) a
PfNF54
EC50 (µM) (SI) a
PfK1
EC50 (µM) (SI) a
Hep G2
CC50 (µM)
11 Molecules 26 01008 i01117.58 ± 5.27 (9.1)ndnd>160
12 Molecules 26 01008 i01229.59 ± 0.93 (0.7)ndnd21.80 ± 1.10
13 Molecules 26 01008 i01323.04 ± 2.00 (>6.9)ndnd>160
14 Molecules 26 01008 i0144.04 ± 0.03 (13.6)ndnd55.17 ± 1.20
15 Molecules 26 01008 i0154.71 ± 0.07 (>8.5)>50>50>40
16 Molecules 26 01008 i01640.42 ± 1.58 (1.9)ndnd76.82 ± 6.50
Mefloquine 2.33 ± 2.290.003 ± 0.0020.117 ± 0.015nd
Chloroquine0.004 ± 0.0010.005 ± 0.0010.212 ± 0.033nd
Artemisininnd0.017 ± 0.0050.011 ± 0.005nd
SDS (0.1%)ndndndb 1%
a Selectivity index (CC50/EC50); b Percent viability; nd = not determined.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nyamwihura, R.J.; Zhang, H.; Collins, J.T.; Crown, O.; Ogungbe, I.V. Nopol-Based Quinoline Derivatives as Antiplasmodial Agents. Molecules 2021, 26, 1008. https://doi.org/10.3390/molecules26041008

AMA Style

Nyamwihura RJ, Zhang H, Collins JT, Crown O, Ogungbe IV. Nopol-Based Quinoline Derivatives as Antiplasmodial Agents. Molecules. 2021; 26(4):1008. https://doi.org/10.3390/molecules26041008

Chicago/Turabian Style

Nyamwihura, Rogers J., Huaisheng Zhang, Jasmine T. Collins, Olamide Crown, and Ifedayo Victor Ogungbe. 2021. "Nopol-Based Quinoline Derivatives as Antiplasmodial Agents" Molecules 26, no. 4: 1008. https://doi.org/10.3390/molecules26041008

Article Metrics

Back to TopTop