Next Article in Journal
Mineral Composition and Graphitization Structure Characteristics of Contact Thermally Altered Coal
Next Article in Special Issue
A Fluoroponytailed NHC–Silver Complex Formed from Vinylimidazolium/AgNO3 under Aqueous–Ammoniacal Conditions
Previous Article in Journal
Stability, Morphology, and Effects of In Vitro Digestion on the Antioxidant Properties of Polyphenol Inclusion Complexes with β-Cyclodextrin
Previous Article in Special Issue
Generation and Reactions of ε-Carbonyl Cations via Group 13 Catalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design and Synthesis of C-1 Methoxycarbonyl Derivative of Narciclasine and Its Biological Activity

1
Department of Chemistry, Brock University, 1812 Sir Isaac Brock Way, St. Catharines, ON L2S 3A1, Canada
2
Department of Chemistry and Biochemistry, Texas State University, San Marcos, TX 78666, USA
*
Authors to whom correspondence should be addressed.
Dedicated to the memory of Victor Snieckus and his legendary contributions to the art and craft of organic synthesis, and to Tomas Hudlicky who was our colleague, mentor, and friend.
Submission received: 30 May 2022 / Revised: 8 June 2022 / Accepted: 9 June 2022 / Published: 14 June 2022

Abstract

:
A 15-step chemoenzymatic total synthesis of C-1 methoxycarbonyl narciclasine (10) was accomplished. The synthesis began with the toluene dioxygenase-mediated dihydroxylation of ortho-dibromobenzene to provide the corresponding cis-dihydrodiol (12) as a single enantiomer. Further key steps included a nitroso Diels–Alder reaction and an intramolecular Heck cyclization. The C-1 homolog 10 was tested and evaluated for antiproliferative activity against natural narciclasine (1) as the positive control. Experimental and spectral data are reported for all novel compounds.

1. Introduction

Narciclasine (1) and pancratistatin (2), Figure 1, are two alkaloids found in the Amaryllidaceae family of plants that have garnered significant scientific attention due to their attractive biological activities. The antineoplastic properties of (1) and (2) have led to intense investigations towards their synthesis, and their scarce availability and poor solubility profiles have also led to synthetic efforts related to their unnatural and/or truncated analogs. Thus, a common objective of these endeavors is to synthesize products with comparable or enhanced anti-tumor activities to those of (1) and (2) with improved bioavailability profiles. In this paper we present our most recent efforts to synthesize such novel compounds, namely the C-1 methoxycarbonyl analog of narciclasine (10), and report findings on its biological activity.
Previous studies [1] of the pharmacophore of Amaryllidaceae alkaloids suggest that structural alterations in the “northwest bay region”, encompassing the region between C-1 and C-10, can be introduced without detriment to the biological activities. The C-1 benzoate derivative of pancratistatin (3) reported by Pettit [2] and the C-1 benzoyloxymethyl compound (4) reported by us [3,4] are the only synthetic derivatives of pancratistatin that exhibited enhanced biological activity when tested (both exhibiting nanomolar activity) [5,6,7,8,9]. These findings suggest that further studies of both pancratistatin and narciclasine analogs may lead to the discovery of novel compounds that exhibit similar or enhanced activities. Pancratistatin is more thoroughly discussed in the literature, even though narciclasine was shown to have the stronger biological activity [10]. The number of unnatural derivatives of narciclasine reported in the literature is very small compared with those derived from pancratistatin. Thus, our group previously published efforts towards the total syntheses of narciclasine analogs, namely compounds 59 shown in Figure 1 [11,12,13,14]. 10-Aza-narciclasine (7) is the only derivative of narciclasine synthesized by our group that displayed activity comparable to that of the natural product [12].
Recently, we began pursuing the synthesis of C-1 narciclasine analogs based on a convergent synthesis wherein the A- and C-ring precursors are synthesized separately and later combined to form the B-ring, thus producing the full narciclasine backbone (Scheme 1). The synthesis of the A-ring precursor (14) begins from ortho-vanillin (15) and is based on previous work [15,16,17,18], while the C-ring synthesis begins with ortho-dibromobenzene (13). Enzymatic dihydroxylation of the latter compound installs the syn-diol moiety at C-3 and C-4 in the correct configuration (12, Scheme 1) [19]. Key steps in the synthesis include a nitroso Diels–Alder reaction that installs the heteroatoms at C-2 and C-5 in the correct stereochemical configuration, and an intramolecular Heck cyclization that forms the C10a-C10b bond. In this paper we report the first total synthesis of a C-1 narciclasine homolog, as well as initial biological evaluation studies for the synthesized compound.

2. Results and Discussion

2.1. Synthesis of the A-Ring Coupling Precursor

The synthesis of the A-ring coupling precursor of narciclasine is based on previous Amaryllidaceae chemistry [15,16] that involves the derivatization of ortho-vanillin (15) to the bromoarene (18) in three steps (Scheme 2). This is followed by a Rieche formylation reaction [17] that installs an aldehyde moiety at C-6a (19) that can be further elaborated to produce the acid chloride 14 in two additional steps [18]. At this stage, the crude acid chloride 14 is coupled to the primary amine of the C-ring precursor (11) to form the necessary amide bond of the B-ring.

2.2. Enzymatic Dihydroxylation of ortho-Dibromobenzene and Nitroso Diels–Alder Reaction

The more elaborate synthesis of the C-ring coupling precursor, which contains all the necessary chiral centers of the final product, begins with the enzymatic dihydroxylation of ortho-dibromobenzene (13, Scheme 3). This is accomplished by a whole-cell fermentation process [20] with a recombinant strain of E. coli JM109 (pDTG601A [21]) that over-expresses toluene dioxygenase and was successfully used for the oxidative de-aromatization of various aromatic substrates [22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37] employed in the synthesis of natural products. Once dibromodiene diol 12 was obtained, it was subjected to acetonide protection and a nitroso Diels–Alder reaction in a one-pot procedure to form the bicyclic oxazine 21. The facial selectivity that is observed in the produced oxazine is attributed to the presence of the acetonide group at C-3 and C-4 that effectively blocks the dienophile approach from the β-face. Additionally, the observed regioselectivity of the cycloaddition reaction stems from the C-Br bond at C-2 that results in a highly polarized diene.

2.3. C–1 Functionalization via Pd-Catalyzed Carbonylation

Upon reduction of the N-O bond of the nitroso Diels–Alder adduct 21 (Scheme 4), conduramine 22 is obtained which contains all the necessary carbon–heteroatom bonds of the C-ring in the correct stereochemical configurations. This operation was performed by treating 21 with aluminum amalgam which simultaneously reduced the N-O bond as well as the C-Br bond at C-2. The product is obtained in modest yields, due to the overreduction of the C-Br bond at C-1 as an unavoidable competing reaction.
After the TBS protection of the C-2 hydroxyl group, the next step is functionalization at C-1. Initial studies concentrated on the conversion of the vinyl bromide moiety of 23 to a vinyl nitrile moiety (25, Scheme 5), with the end-goal being reduction of the nitrile to an aldehyde group. This was done because installation of an aldehyde at C-1 has the potential for further derivatization to both hydroxymethyl and aminomethyl derivatives. While the installation of a nitrile group at C-1 was performed successfully, treatment of the resulting nitrile 25 with a hydride source (for example, DIBAL-H) resulted in the reduction of the nitrile with concomitant loss of the C-2 silyl ether group, presumably via a 1,4-conjugate elimination reaction (Scheme 5). Thus, a different route was pursued based on previous work [38] from our group in which a Pd-catalyzed carbonylation reaction was carried out in the presence of carbon monoxide and methanol to produce the corresponding methyl ester (24, Scheme 4).
The final operation that was performed on this coupling precursor is the deprotection of the benzyl carbamate to reveal a primary amine (11, Scheme 4) that can be condensed with the A-ring coupling partner, acid chloride 14. This could not be carried out under classical hydrogenolysis [39] or hydrolysis conditions because of the presence of incompatible functional groups, namely the C10b-C1 olefin and the C-1 methyl ester. However, we were pleased to learn that this transformation can be successfully carried out by subjecting the substrate to a transfer hydrogenation reaction [40] by treatment of 24 with triethylsilane and palladium chloride to furnish amine 11.

2.4. Intramolecular Heck Reaction and Global Deprotection to Afford 1-Methoxycarbonyl Narciclasine

Following the deprotection of the benzyl carbamate, the resulting amine (11) was subjected to a condensation reaction with acid chloride 14 to produce adduct 27 (Scheme 6) which contains the full narciclasine backbone, except for the C10a–C10b bond. Following a Boc protection reaction to form imide 28, the key intramolecular Heck reaction was performed to establish this bond. This transformation was successfully accomplished to afford the cyclized product (29) by employing reaction conditions that were previously used by our group in the synthesis of aza-narciclasine derivatives 5, 6 and 7. The final stage of this synthesis is the global deprotection of 29 to furnish the desired product in two more steps. We again referred to previously established protocols that our group used in the synthesis of pancratistatin and narciclasine analogs [11,12]. Demethylation was carried out first (LiCl, DMF, 90 °C) and under these conditions thermal N-Boc deprotection also occurred to obtain 30. Desilylation and acetonide deprotection were successfully accomplished by treating 30 with HCl to obtain the target product, 1–methoxycarbonyl narciclasine (10).

2.5. Biological Activity

We used an in vitro cancer model—lung adenocarcinoma cell line A549 (Figure 2)—to evaluate the ability of compound 10 to inhibit the proliferation of cancer cells. Narciclasine was used as a control. As expected, the narciclasine control showed good nanomolar activity against this cell line, displaying an IC50 of 20 nM. Compound 10 was much less potent, yielding an IC50 of 15.5 μM (Figure 2). Thus, it appears that derivatization of narciclasine at C-1 with the methoxycarbonyl group leads to a reduction in activity although it does not abolish it completely.

3. Materials and Methods

3.1. General Information

All solvents were distilled and kept dry before usage. All reactions were done in inert atmosphere (Ar or N2) and at room temperature, unless otherwise stated. All reagents were obtained from commercial sources. Nuclear magnetic resonance (NMR) analyses were performed on Bruker Avance AV 300, Bruker Avance III HD 400 and Bruker Avance AV 600 digital NMR spectrometers, running Topspin 2.1 and 3.5 software. The probes are furnished with VT (variable temperature) and gradient equipment. Chemical shifts are given in δ, relative integral, multiplicity (singlet (s), doublet (d), triplet (t), quartet (q), multiplet (m)) and coupling constants (J) in Hz. Melting points (m.p.) were measured in a capillary apparatus. Mass spectra (HRMS) measurements were determined on an LTQ Orbitrap XL. The molecular mass-associated ion was measured by electron ionization, electrospray ionization or fast atom bombardment. Infrared (IR) spectra were recorded on an FT-IR spectrophotometer as neat and are reported in wave numbers (cm−1) and intensity (broad (br), strong (s), medium (m), weak (w)). Column chromatography was performed on flash grade 60 silica gel. Thin-layer chromatography (TLC) was performed on silica gel 60 F254-coated aluminum sheets. TLC plates were visualized using UV and stained with iodine, cerium ammonium molybdate (CAM), KMnO4 solutions, FeCl3 solutions, ninhydrin solutions or 2,4-dinitrophenylhydrazine (2,4-DNP) solutions.
The five-step synthesis of compound 20 from ortho-vanillin (15) was accomplished by repetition of known methods [15,16,17,18]. Experimental and spectral data are reported for all novel compounds (see Supporting Materials).

3.2. Preparation and Characterization of Compounds

3.2.1. (3aS,4S,7R,7aS)-Benzyl-4,5-dibromo-2,2-dimethyl-3a,4,7,7a-tetrahydro-4,7-(epoxyi-mino)benzo[d][1,3]dioxole-8-carboxylate (21)

To the solution of dibromo dienediol 12 [19,20] (1.68 g, 6.2 mmol) in 2,2-DMP (40 mL) was added pTsOH (107 mg, 0.6 mmol). This mixture was allowed to stir at room temperature for approximately 2 h until a TLC analysis showed complete consumption of the starting material. At this point, NaIO4 (2.65 g, 12.4 mmol) and H2O (10 mL) were added to the reaction mixture and the contents were cooled to 0 °C. A separate solution of benzyl hydroxycarbamate (CbzNHOH, 1.55 g, 9.3 mmol) in MeOH (20 mL) was prepared and added dropwise to the reaction mixture over a period of 30 min. The reaction mixture was left to warm up to room temperature and stirred for 16 h at room temperature, then it was quenched with a saturated NaHSO3 solution (20 mL) and the volume of the mixture was reduced by rotary evaporation. The aqueous layer was extracted with EtOAc (3 × 40 mL), the organic layers were combined and dried with Na2SO4, filtered, and the filtrate was concentrated by rotary evaporation. The crude product was purified by column chromatography [silica gel, EtOAc:Hex (1:3)] to obtain 21 as a colorless gel (1.9 g, 74 %).
21: Rf = 0.6 [EtOAc:Hex (1:3)]; [ α ] D 23 = −5.4 (c = 0.36, CHCl3); IR (neat, cm−1) 3030, 2997, 1755, 1725, 1598, 1442, 1377, 1268, 1225, 1074; 1H NMR (300 MHz, CDCl3) δ 7.39–7.30 (m, 5H), 6.69 (d, J = 6.1 Hz, 1H), 5.21 (s, 2H), 5.08 (dd, J = 6.4, 4.1 Hz, 1H), 4.79 (d, J = 6.9 Hz, 1H), 4.56 (dd, J = 6.9, 4.0 Hz, 1H), 1.39 (s, 3H), 1.33 (s, 3H); 13C NMR (75 MHz, CDCl3) δ 157.3, 135.2, 131.3, 128.6 (2C), 128.5, 127.9 (2C), 122.1, 111.9, 93.8, 82.3, 73.8, 68.7, 55.4, 25.8, 25.5. HRMS (EI) calcd for C16H14Br2NO5[M-CH3]+: 457.9233. Found 457.9232; Anal. Calcd for C17H17Br2NO5: C, 42.97; H, 3.61. Found C, 43.19; H, 3.85.

3.2.2. ((3aS,4R,7R,7aR)-Benzyl-6-bromo-7-hydroxy-2,2-dimethyl-3a,4,7,7a-tetrahydroben-zo[d][1,3]dioxol-4-yl)carbamate (22)

Aluminum amalgam (prepared from 20 mg, 0.72 mmol, 8 equivalents of aluminum turnings after treatment with 1 M aq. KOH and 0.5% aq. HgCl2) was added to a stirred solution of the oxazine 21 (43 mg, 0.09 mmol) in aqueous THF (THF:H2O, 10:1, 1 mL), cooled to 0 °C, over 3 days (2 equivalents at reaction start time, then another 2 equivalents every 24 h). After each addition, the reaction was allowed to attain room temperature and stirring was continued until the next addition. After about 96 h, TLC analysis confirmed reaction completion. The reaction mixture was diluted with 2 mL of MeOH, stirred for 10 min, then filtered through Celite. The organic layer was dried with Na2SO4 and concentrated. The residue was purified by column chromatography [silica gel, EtOAc:Hex (1:2)] to afford the alcohol 22 as a colorless oil (23 mg, 40%).
22: Rf = 0.5 [EtOAc:Hex (1:2)]; [ α ] D 23 = −20.7 (c = 0.46, CHCl3); IR (neat, cm−1) 3334, 2933, 1697, 1516, 1374, 1240, 1212, 1035; 1H NMR (300 MHz, CDCl3) δ 7.34 (s, 5H), 6.33 (d, J = 6.1 Hz, 1H), 5.49 (d, J = 9.0 Hz, 1H), 5.09 (s, 2H), 4.50 (dd, J = 6.7, 2.5 Hz, 1H), 4.43–4.30 (m, 3H), 3.41 (s, 1H), 1.40 (s, 3H), 1.31 (s, 3H); 13C NMR (75 MHz, CDCl3) δ 155.8, 136.2, 130.3, 128.7, 128.4 (2C), 128.3 (2C), 126.9, 108.9, 79.0, 75.9, 73.1, 67.3, 50.9, 26.7, 24.6. HRMS (EI) calcd for C16H17BrNO5[M-CH3]+: 382.0285. Found 382.0280; Anal. Calcd for C17H20BrNO5: C, 51.27; H, 5.06. Found C, 51.50; H, 5.11.

3.2.3. ((3aS,4R,7R,7aS)-Benzyl-6-bromo-7-((tert-butyldimethylsilyl)oxy)-2,2-dimethyl-3a,4,7,7a-tetrahydrobenzo[d][1,3]dioxol-4-yl)carbamate (23)

Imidazole (136 mg, 2.0 mmol) and TBSCl (452 mg, 3.0 mmol) were added to a solution of alcohol 22 (402 mg, 1.0 mmol) in 10 mL of DCM at room temperature. After the complete consumption of the starting material was observed by TLC (approximately 24 h), the reaction mixture was quenched with a saturated aqueous solution of NH4Cl (10 mL) and extracted with DCM (3 × 10 mL). The combined organic phases were dried over Na2SO4 and concentrated. The residual oil was purified by column chromatography [silica gel, EtOAc:Hex (1:4)] to afford the product 23 as a pale-yellow oil (439 mg, 86%).
23: Rf = 0.6 [EtOAc:Hex (1:4)]; [ α ] D 22 = −4.5 (c = 0.30, CHCl3); IR (neat, cm−1) 3398, 2931, 1720, 1637, 1512, 1384, 1224, 1065; 1H NMR (300 MHz, CDCl3) δ 7.35 (s, 5H), δ 6.50 (d, J = 6.5 Hz, 1H), δ 5.69 (d, J = 5.69 Hz, 1H), δ 5.11 (s, 1H), δ 4.55 (s, 1H), δ 4.36 (s, 1H), δ 1.40 (s, 3H), δ 1.32 (s, 3H), δ 0.91 (s, 9H), δ 0.22 (s, 6H); 13C NMR (75 MHz, CDCl3) δ 155.7, 136.6, 131.6, 128.7, 128.3, 127.4, 109.0, 80.1, 74.7, 67.1, 49.6, 30.0, 26.6, 25.9, 24.8, 18.1, −4.1, −4.8. HRMS (EI) calcd for C22H31BrNO5Si[M-CH3]+: 496.1149. Found 496.1149. Anal. Calcd for C23H34BrNO5Si: C, 53.90; H, 6.69. Found C, 54.00; H, 6.89.

3.2.4. (3aS,4S,7R,7aS)-Methyl 7-(((benzyloxy)carbonyl)amino)-4-((tert-butyldimethylsilyl) oxy)-2,2-dimethyl-3a,4,7,7a-tetrahydrobenzo[d][1,3]dioxole-5-carboxylate (24)

Pd(PPh3)4 (81 mg, 0.07 mmol) and distilled Et3N (0.08 mL, 0.54 mmol) were added to an anhydrous stirred solution of the vinyl bromide 23 (137 mg, 0.27 mmol) in DMF:MeOH (1:1, 4 mL). The mixture was degassed with CO gas for 15 min and then a CO balloon was fixed to the top of the flask and it was heated to 65 °C for 24 h. Once complete consumption of the starting material was observed by TLC analysis, the reaction mixture was allowed to cool to room temperature and then diluted with DCM and filtered through a plug of Celite. The resulting filtrate was extracted with a saturated solution of NH4Cl (5.0 mL) and washed with DCM (3 × 10 mL). The combined organic layers were dried over Na2SO4 and concentrated. The residue was purified by column chromatography [silica gel, EtOAc:Hex (1:4)] to obtain the product 24 as a white solid (72 mg, 54%).
24: Rf = 0.5 [silica gel, EtOAc:Hex (1:4)]; mp 77 °C (EtOAc); [ α ] D 22 = −12.3 (c = 0.10, CHCl3); IR (neat, cm−1) 3383, 3066, 3033, 2952, 2930, 2900, 2857, 1717, 1504; 1H NMR (300 MHz, CDCl3) δ 7.39–7.27 (m, 5H), 5.77 (d, J = 10.3 Hz, 1H), 5.14–5.01 (m, 2H), 4.79 (s, 1H), 4.75–4.64 (m, 1H), 4.48 (dt, J = 6.8, 4.3 Hz, 2H), 3.79 (s, 3H), 1.28 (d, J = 2.5 Hz, 3H), 1.27 (s, 3H), 0.83 (s, 9H), 0.18 (s, 3H), 0.06 (s, 3H). 13C NMR (100 MHz, CDCl3) δ 166.1, 155.8, 140.9, 136.5, 135.6, 128.7, 128.3, 108.6, 88.0, 78.6, 77.8, 77.4, 76.9, 72.7, 67.2, 64.9, 52.4, 47.2, 26.4, 26.0, 24.4, 18.2, −4.5, −4.9. HRMS (EI) calcd for C25H37NO7Si: 491.2334. Found 491.2338.

3.2.5. (3aS,4S,7R,7aS)-Methyl-7-(6-bromo-4-methoxybenzo[d][1,3]dioxole-5-carboxamido) -4-((tert-butyldimethylsilyl)oxy)-2,2-dimethyl-3a,4,7,7a-tetrahydrobenzo[d][1,3]dioxole-5 -carboxylate (27)

  • Step A: Preparation of acid chloride 14
A flame-dried flask equipped with a stir bar was charged with the acid 20 (509 mg, 1.85 mmol). After the flask contents were dissolved in DCM (16 mL), (COCl)2 was added (0.32 mL, 3.7 mmol). A catalytic amount of DMF was then added (approximately 5 drops) to the solution, and the reaction mixture was allowed to stir at room temperature for 3 h. The volume of the reaction mixture was then reduced under a rotary evaporator [which was equipped with a bubbler containing a 10% NaOH (aq.) solution] and the crude acid chloride 14 was then dried further under high vacuum. It was immediately used in step C without further purification.
  • Step B: Preparation of primary amine 11
A flame-dried flask equipped with a stir bar was charged with the substrate 24 (455 mg, 0.93 mmol) and PdCl2 (181 mg, 1.02 mmol). After the contents were dissolved in DCM (22 mL), Et3N (0.39 mL, 2.8 mmol) and Et3SiH (0.45 mL, 2.8 mmol) were added, respectively. After TLC analysis confirmed the complete consumption of the starting material (30 min), the reaction mixture was filtered through a plug of Celite and concentrated to dryness under high vacuum. It was immediately used in step C without further purification.
  • Step C: Condensation of amine 11 with acid chloride 14 to afford compound 27
The crude amine 11 (0.93 mmol) and DMAP (11 mg, 0.09 mmol) were charged to flame-dried flask equipped with a stir bar. The reaction mixture contents were then dissolved in DCM (14.0 mL) and pyridine (0.23 mL, 2.8 mmol) was added. The reaction mixture was then cooled to 0 °C. A separate solution of the crude acid chloride 14 (1.85 mmol) in DCM (8.0 mL) was prepared and added to the reaction mixture over 10 min. The reaction mixture was then allowed to slowly warm to room temperature over 1 h and then stirred for another 12 h. Next, the reaction mixture was concentrated and dried under high vacuum, and the residue purified by column chromatography [silica gel, EtOAc:Hex (1:1)] to obtain the product 27 as a pale-yellow oil (411 mg, 72%).
27: Rf = 0.6 [silica gel, EtOAc:Hex (1:1)]; [ α ] D 22 = −28.0 (c = 0.30, CHCl3). IR (neat, cm−1) 3369, 2927, 2855, 1721, 1679,1504, 1468; 1H NMR (300 MHz, CDCl3) δ 6.70 (s, 1H), 6.64 (d, J = 10.3 Hz, 1H), 6.02 (s, 1H), 5.96 (s, 2H), 5.26 (dd, J = 14.6, 4.5 Hz, 1H), 4.83 (s, 1H), 4.66 (d, J = 6.9 Hz, 1H), 4.51 (d, J = 4.7 Hz, 1H), 3.99 (s, 3H), 3.82 (s, 3H), 1.33 (s, 3H), 1.31 (s, 3H), 0.63 (s, 7H), 0.14 (s, 3H), 0.01 (s, 3H). 13C NMR (150 MHz, CDCl3) δ 165.9, 164.8, 150.4, 141.2, 140.4, 136.0, 135.8, 125.5, 111.3, 108.5, 107.2, 101.8, 78.4, 76.5, 64.9, 60.4, 52.2, 45.4, 26.2, 25.3, 24.2, 17.5, −4.9, −5.0. HRMS (EI) Calcd for C26H36BrNO9Si: 613.1337, Found 613.1328.

3.2.6. (3aS,4S,7R,7aS)-Methyl-7-(6-bromo-N-(tert-butoxycarbonyl)-4-methoxybenzo[d][1,3] dioxole-5-carboxamido)-4-((tert-butyldimethylsilyl)oxy)-2,2-dimethyl-3a,4,7,7a-tetrahyd-robenzo[d][1,3]dioxole-5-carboxylate (28)

The amide 27 (388 mg, 0.63 mmol) and DMAP (158 mg, 1.29 mmol) were charged to a flame-dried RBF with a stir bar, then dissolved in dry MeCN (12 mL). Next, Boc2O (275 mg, 1.26 mmol) was added. The reaction was left to stir at room temperature for 4 h before quenching with H2O (20 mL) and extracting with DCM (3 × 20 mL). The organic layers were then combined, dried with MgSO4 and concentrated to dryness under reduced pressure. The residue was purified by column chromatography [silica gel, EtOAc:Hex (1:2)] to obtain 28 (338 mg, 75%) as a yellow oil.
28: Rf = 0.6 [silica gel, EtOAc:Hex (1:2)]. [ α ] D 22 = −60.4 (c = 0.45, CHCl3). IR (neat, cm−1) 2986, 2956, 2934, 2859, 1729, 1679, 1623. 1H NMR (mixture of rotamers) (300 MHz, CDCl3) δ 6.72 (d, J = 3 Hz, 1H), 6.62 (s, 1H), 5.97 (s, 2H), 5.52 (s, 1H), 4.78 (s, 1H), 4.57 (m, 1H), 4.30 (m, 1H), 3.97 (s, 3H), 3.74 (s, 3H), 1.52 (m, 3H), 1.39 (m, 3H), 1.33 (m, 9H), 0.87 (s, 9H), 0.20 (s, 3H), 0.13 (s, 3H). 13C NMR (150 MHz, CDCl3) δ 166.5, 151.5, 150.1, 148.0, 137.7, 130.7, 108.6, 107.4, 107.1, 102.0, 84.4, 82.2, 80.3, 76.7, 66.2, 64.0, 62.5, 60.4, 60.3, 54.5, 51.8, 27.8, 26.4, 25.9, 18.1, −4.6, −5.0. HRMS (EI) Calcd for C31H44BrNO11Si: 713.1862, Found 713.1878.

3.2.7. 4-(tert-Butyl)-11-methyl(3aS,3bR,12S,12aS)-12-((tert-butyldimethylsilyl)oxy)-6-met-hoxy-2,2-dimethyl-5-oxo-3a,5,12,12a-tetrahydrobis([1,3]dioxolo)[4,5-c:4’,5’-j]phenanthrid-ine-4,11(3bH)-dicarboxylate (29)

A flame-dried flask with a stir bar was charged with the substrate 28 (123 mg, 0.17 mmol), Pd(OAc)2 (10 mg, 0.04 mmol) and Ag3PO4 (214 mg, 0.51 mmol). The reaction mixture contents were then dissolved in dry toluene (9 mL) and degassed with argon for 15 min. Next, 1,2–bis(diphenylphosphino)ethane (14 mg, 0.03 mmol) was added and the reaction mixture was heated to 130 °C for 16 h. After cooling to room temperature, the reaction mixture was filtered through a plug of Celite and then concentrated under reduced pressure. The residue was purified by column chromatography [silica gel, EtOAc:Hex (1:2)] to obtain 29 (44 mg, 40%, 61% brsm) as a colorless oil and 41 mg of recovered starting material 28.
29: Rf = 0.4 [silica gel, EtOAc:Hex (1:2)]. [ α ] D 22 = 130.0 (c = 0.8, CHCl3). IR (neat, cm−1) 2989, 2956, 2934, 2857, 1722, 1666, 1606. 1H NMR (400 MHz, CDCl3) δ 6.71 (s, 1H), 6.03 (d, J = 1.4 Hz, 2H), 4.89 (d, J = 6.1 Hz, 1H), 4.58–4.54 (m, 1H), 4.08–4.04 (m, 2H), 4.03 (s, 3H), 3.76 (s, 3H), 1.53 (s, 3H), 1.52 (s, 9H), 1.31 (s, 3H), 0.90 (s, 9H), 0.17 (s, 3H), 0.14 (s, 3H). 13C NMR (100 MHz, CDCl3) δ 168.1, 161.4, 152.7, 152.1, 144.4, 139.6, 133.6, 129.7, 126.6, 118.3, 111.5, 102.4, 100.2, 83.1, 79.4, 79.0, 73.4, 61.4, 56.0, 52.5, 28.1(3C), 27.2, 25.8 (3), 25.1, 18.2, −4.3, −5.1. HRMS (EI) calcd for C27H34NO11Si [M-C4H9]+: 576.1896. Found 576.1905.

3.2.8. (3aS,3bR,12S,12aS)-Methyl-12-((tert-butyldimethylsilyl)oxy)-6-hydroxy-2,2-dimet-hyl-5-oxo-3a,3b,4,5,12,12a-hexahydrobis([1,3]dioxolo)[4,5-c:4’,5’-j]phenanthridine-11-car-boxylate (30)

LiCl (28 mg, 0.66 mmol) was added to a solution of the substrate 29 (28 mg, 0.044 mmol) in dry DMF (3 mL). The reaction mixture contents were then subjected to three cycles of freeze-pump-thaw, before the mixture was heated to 90 °C for 12 h. The reaction mixture was then cooled to room temperature and filtered through a plug of Celite, then concentrated under reduced pressure. The residue was purified by column chromatography [silica gel, EtOAc:Hex (1:2)] to obtain 30 (18 mg, 80%) as a white solid.
30: Rf = 0.6 [silica gel, EtOAc:Hex (1:2)]; mp 260 °C (decomp., EtOAc); [ α ] D 22 = 38.4 (c = 0.85, CHCl3). IR (neat, cm−1) 3205, 3181, 3088, 3034, 2997, 2954, 2929, 2857, 1731, 1672, 1620, 1599. 1H NMR (600 MHz, CDCl3) δ 13.06 (s, 1H), 6.57 (s, 1H), 6.31 (s, 1H), 6.05 (s, 2H), 4.49 (d, J = 1 Hz, 1H), 4.17 (dd, J = 2, 1 Hz, 1H), 4.08 (dd, J = 2, 1 Hz, 1H), 4.03 (d, J = 1 Hz, 1H), 3.82 (s, 3H), 1.51 (s, 3H), 1.36 (s, 3H), 0.91 (s, 9H), 0.18 (s, 3H), 0.14 (s, 3H). 13C NMR (150 MHz, CDCl3) δ 168.4, 167.4, 152.9, 146.7, 135.4, 132.2, 126.9, 124.2, 111.5, 106.2, 102.7, 97.5, 79.1, 78.2, 73.5, 54.9, 52.7, 27.3, 25.8, 25.1, 18.1, −4.3, −5.2. HRMS (EI) calcd for C25H33NO9Si: 519.1919. Found 519.1923.

3.2.9. (2S,3R,4S,4aR)-Methyl-2,3,4,7-tetrahydroxy-6-oxo-2,3,4,4a,5,6-hexahydro-[1,3]diox-olo [4,5-j]phenanthridine-1-carboxylate (10)

HCl (0.1 mL, 1.1M) was added to a solution of the substrate 30 (16 mg, 0.031 mmol) in THF (1 mL) that was cooled to 0 °C. The reaction mixture was allowed to attain room temperature over 1 h and then stirred at room temperature for a further 36 h. The reaction mixture was then extracted with EtOAc (5 mL × 5) and the combined organic layers were dried with MgSO4, then concentrated under reduced pressure. The residue was purified by column chromatography [silica gel, MeOH:DCM (1:5)] to obtain 10 (6 mg, 55%) as a white solid.
10: Rf = 0.6 [silica gel, MeOH:DCM (1:5)]; mp > 300 °C (decomp., EtOAc); [ α ] D 20   = 141.5 (c = 0.30, MeOH); IR [neat, cm−1] 3666, 3297, 3099, 2983, 2907, 2796, 2669, 2119, 1698, 1672, 1595, 1508; 1H NMR [400 MHz, (CD3)2CO] δ 12.91 (s, 1H), 7.39 (s, 1H), 6.29 (s, 1H), 6.09 (dd, J = 7.5, 1.1 Hz, 2H), 4.78 (s, 2H), 4.65 (d, J = 3.0 Hz, 1H), 4.42–4.35 (m, 1H), 4.05 (t, J = 3.0 Hz, 1H), 4.04 (s, 1H), 3.64 (s, 3H); 13C NMR [100 MHz, (CD3)2CO] δ 170.8, 169.9, 152.9, 146.5, 135.8, 132.6, 131.1, 130.3, 108.1, 103.5, 99.3, 73.7, 72.8, 70.3, 54.7, 52.4; HRMS (EI) calcd. for C16H15NO9: 365.0741. Found 365.0749.

3.3. Cell Viability Assay

Cell viability was measured by MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide) assay in lung adenocarcinoma cell line A549 (obtained from American Type Culture Collection). A549 cells were maintained in RPMI-1640 media supplemented with 10% FBS. For measuring the IC50 values of cell viability, 3000 cells per well were plated into a 96-well plate, and the cells were treated with a series of dilutions (from 500 µM to 0.32 nM) of individual compounds in triplicate for four days. After four days, the cells were replaced with MTT reagent (at 0.25 mg/mL in culture media) in each well and incubated with cells for 2 h at 37 °C. The culture media was removed after the microplates were spun at 2000 rpm for 5 min. DMSO was then used to dissolve the crystals formed in each of the 96 wells. Optical density values at wavelength 570 nm and 630 nm were measured using SpectraMax 190 (Molecular Devices, San Jose, CA, USA), and the difference in the two optical density values was used to analyze the relative cell survival in each well. IC50 values were calculated using Graphpad Prism software.

4. Conclusions

The first C-1 homolog of narciclasine, methoxycarbonyl narciclasine (10), was prepared in fifteen steps via a chemoenzymatic total synthesis from ortho-dibromodiene diol (12) and ortho-vanillin (15). The enhanced biological activities of C-1 homologs of pancratistatin (such as compounds 3 and 4 [2,3]) provided the incentive to pursue the analogous synthesis of C-1 homologs of narciclasine. Our synthesis and biological testing of compounds 9, described elsewhere, and 10, described here, show that derivatization of narciclasine at position C-1 leads to reduced antiproliferative potencies, although it does not abolish activity completely. This stands in contrast to similar efforts in pancratistatin area where more potent analogs can be achieved through C-1 modification [2,3]. Current efforts are underway to synthesize other unnatural C-1 narciclasine homologs.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/molecules27123809/s1; 1H and 13C NMR spectral data of compounds 10, 21–24 and 27–30.

Author Contributions

T.H. and L.H. conceived and designed the experiments; L.H. performed the experiments and analyzed the data; M.S. and L.D. performed the biological evaluations under the supervision of A.K.; writing—original draft prepared by L.H; writing—review and editing by T.H., K.B. and A.K. All authors have read and agreed to the published version of the manuscript.

Funding

We are grateful to the following agencies for financial support of this work: the Natural Sciences and Engineering Research Council of Canada (NSERC; Idea to Innovation and Discovery grants), the Canada Research Chair Program, the Canada Foundation for Innovation (CFI), TDC Research, Inc., the TDC Research Foundation, the Ontario Partnership for Innovation and Commercialization (OPIC), and The Advanced Biomanufacturing Centre (Brock University). We thank Dr. Liqun Qui for her help with mass spectrometry. A.K. acknowledges the National Institutes of Health (grant 1R15CA227680-01A1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References

  1. McNulty, J.; Larichev, V.; Pandey, S. A synthesis of 3-deoxydihydrocoricidine: Refinement of a structurally minimum pancratistatin pharmacophore. Bioorg. Med. Chem. Lett. 2005, 15, 5315–5318. [Google Scholar] [CrossRef] [PubMed]
  2. Pettit, G.R.; Melody, N.; Herald, D.L. Antineoplastic agents. 450. synthesis of (+)-pancratistatin from (+)-narciclasine as relay. J. Org. Chem. 2001, 66, 2583–2587. [Google Scholar] [CrossRef] [PubMed]
  3. Vshyvenko, S.; Scattolon, J.; Hudlicky, T.; Romero, A.E.; Kornienko, A. Synthesis of C-1 homologues of pancratistatin and their preliminary biological evaluation. Bioorg. Med. Chem. Lett. 2011, 21, 4750–4752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Vshyvenko, S.; Scattolon, J.; Hudlicky, T.; Romero, A.E.; Kornienko, A.; Ma, D.; Tuffley, I.; Pandey, S. Unnatural C-1 homologues of pancratistatin—synthesis and promising biological activities. Can. J. Chem. 2012, 90, 932–943. [Google Scholar] [CrossRef]
  5. Chrétien, F.; Ahmed, S.I.; Masion, A.; Chapleur, Y. Enantiospecific synthesis and biological evaluation of seco analogues of antitumor amaryllidaceae alkaloids. Tetrahedron 1993, 49, 7463–7478. [Google Scholar] [CrossRef]
  6. McNulty, J.; Mao, J.; Gibe, R.; Mo, R.; Wolf, S.; Pettit, G.R.; Herald, D.L.; Boyd, M.R. Studies directed towards the refinement of the pancratistatin cytotoxic pharmacophore. Bioorg. Med. Chem. Lett. 2001, 11, 169–172. [Google Scholar] [CrossRef]
  7. Phung, A.N.; Zannetti, M.T.; Whited, G.; Fessner, W.D. Stereospecific biocatalytic synthesis of pancratistatin analogues. Angew. Chemie Int. Ed. 2003, 42, 4821–4824. [Google Scholar] [CrossRef]
  8. Ibn-Ahmed, S.; Khaldi, M.; Chrétien, F.; Chapleur, Y. A short route to enantiomerically pure benzophenanthridinone skeleton:  synthesis of lactone analogues of narciclasine and lycoricidine. J. Org. Chem. 2004, 69, 6722–6731. [Google Scholar] [CrossRef]
  9. Hudlicky, T.; Rinner, U.; Finn, K.J.; Ghiviriga, I. Reactions of indole derivatives with oxiranes and aziridines on silica. Synthesis of β-carbolin-1-one mimic of pancratistatin. J. Org. Chem. 2005, 70, 3490–3499. [Google Scholar] [CrossRef] [PubMed]
  10. Rinner, U.; Hillebrenner, H.L.; Adams, D.R.; Hudlicky, T.; Pettit, G.R. Synthesis and biological activity of some structural modifications of pancratistatin. Bioorg. Med. Chem. Lett. 2004, 14, 2911–2915. [Google Scholar] [CrossRef]
  11. Vshyvenko, S.; Reisenauer, M.R.; Rogelj, S.; Hudlicky, T. Synthesis and biological evaluation of unnatural derivatives of narciclasine: 7-aza-nornarciclasine and its N-oxide. Bioorg. Med. Chem. Lett. 2014, 24, 4236–4238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Vshyvenko, S.; W’Giorgis, Z.; Weber, A.; Neverova, N.; Hedberg, B.; Hudlicky, T. Synthesis and biological activity of 10-aza-narciclasine. Adv. Synth. Catal. 2015, 357, 83–87. [Google Scholar] [CrossRef]
  13. Ticli, V.; Zhao, Z.; Du, L.; Kornienko, A.; Hudlicky, T. Synthesis and biological evaluation of 10-benzyloxy-narciclasine. Tetrahedron 2021, 101, 135505. [Google Scholar] [CrossRef] [PubMed]
  14. Stollmaier, J.G.; Thomson, J.; Endoma-Arias, M.A.; Simionescu, R.; Vernaza, A.; Mesa-Diaz, N.; Smith, M.; Du, L.; Kornienko, A.; Hudlicky, T. Conversion of natural narciclasine to its C-1 and C-6 derivatives and their antitumor activity evaluation. Some unusual chemistry of narciclasine. Molecules, 2022; submitted. [Google Scholar]
  15. Magnus, P.; Sebhat, I.K. Application of the β-azidonation reaction to the synthesis of the antitumor alkaloid (+)-pancratistatin. Tetrahedron 1998, 54, 15509–15524. [Google Scholar] [CrossRef]
  16. Schmidt, B.; Riemer, M. Suzuki–Miyaura coupling of halophenols and phenol boronic acids: Systematic investigation of positional isomer effects and conclusions for the synthesis of phytoalexins from pyrinae. J. Org. Chem. 2014, 79, 4104–4118. [Google Scholar] [CrossRef]
  17. Furneaux, R.H.; Gainsford, G.J.; Mason, J.M. Synthesis of the Enantiomers of Hexahydrodibenz[d,f]azecines. J. Org. Chem. 2004, 69, 7665–7671. [Google Scholar] [CrossRef]
  18. Cho, Y.S.; Cho, C.G. Improved total synthesis of (±)-trans-dihydronarciclasine, devised for large-scale preparation. Tetrahedron 2008, 64, 2172–2177. [Google Scholar] [CrossRef]
  19. Finn, K.J.; Collins, J.; Hudlicky, T. Toluene dioxygenase-mediated oxidation of dibromobenzenes. Absolute stereochemistry of new metabolites and synthesis of (−)-conduritol E. Tetrahedron 2006, 62, 7471–7476. [Google Scholar] [CrossRef]
  20. Endoma, M.A.; Bui, V.P.; Hansen, J.; Hudlicky, T. Medium-scale preparation of useful metabolites of aromatic compounds via whole-cell fermentation with recombinant organisms. Org. Process Res. Dev. 2002, 6, 525–532. [Google Scholar] [CrossRef]
  21. Zylstra, G.J.; Gibson, D.T. Toluene degradation by Pseudomonas putida F1. Nucleotide sequence of the todC1C2BADE genes and their expression in Escherichia coli. J. Biol. Chem. 1989, 264, 14940–14946. [Google Scholar] [CrossRef]
  22. Bedard, K.; Hudlicky, T.C. 2-Enzymatic dihydroxylation of aromatic compounds: Nature’s unique reaction and its impact on the synthesis of natural products. In Strategies and Tactics in Organic Synthesis; Michael, H., Ed.; Academic Press: Cambridge, MA, USA, 2021; Volume 15, pp. 53–97. [Google Scholar]
  23. Lan, P.; Ye, S.; Banwell, M.G. The application of dioxygenase-based chemoenzymatic processes to the total synthesis of natural products. Chem. Asian J. 2019, 14, 4001–4012. [Google Scholar] [CrossRef] [PubMed]
  24. Taher, E.S.; Banwell, M.G.; Buckler, J.N.; Yan, Q. The exploitation of enzymatically-derived cis-1,2-dihydrocatechols and related compounds in the synthesis of biologically active natural products. Chem. Rec. 2018, 18, 239–264. [Google Scholar] [CrossRef] [PubMed]
  25. Lewis, S.E. Biotransformations of Arenes: An Overview. In Arene Chemistry; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; pp. 915–937. [Google Scholar]
  26. Reed, J.W.; Hudlicky, T. The quest for a practical synthesis of morphine alkaloids and their derivatives by chemoenzymatic methods. Acc. Chem. Res. 2015, 48, 674–687. [Google Scholar] [CrossRef] [PubMed]
  27. Lewis, S.E. Applications of biocatalytic arene ipso, ortho cis-dihydroxylation in synthesis. Chem. Commun. 2014, 50, 2821–2830. [Google Scholar] [CrossRef] [Green Version]
  28. Hassan, H. Potential pharmaceutical compounds from dioxygenase-derived chiral metabolites. Curr. Trends Biotechnol. Pharm. 2012, 6, 389–397. [Google Scholar]
  29. Banwell, M.G.; Lehmann, A.L.; Menon, R.S.; Willis, A.C. New methods for the synthesis of certain alkaloids and terpenoids. Pure Appl. Chem. 2011, 83, 411–423. [Google Scholar] [CrossRef]
  30. Mortison, J.D.; Sherman, D.H. Frontiers and opportunities in chemoenzymatic synthesis. J. Org. Chem. 2010, 75, 7041–7051. [Google Scholar]
  31. Hudlicky, T. Recent chemoenzymatic total syntheses of natural and unnatural products: Codeine, balanol, pancratistatin, and oseltamivir. Pure Appl. Chem. 2010, 82, 1785–1796. [Google Scholar] [CrossRef]
  32. Hudlicky, T.; Reed, J.W. Celebrating 20 years of SYNLETT-special account on the merits of biocatalysis and the impact of arene cis-dihydrodiols on enantioselective synthesis. Synlett 2009, 2009, 685–703. [Google Scholar] [CrossRef]
  33. Hudlicky, T.; Reed, J.W. Applications of biotransformations and biocatalysis to complexity generation in organic synthesis. Chem. Soc. Rev. 2009, 38, 3117–3132. [Google Scholar] [CrossRef]
  34. Boyd, D.R.; Bugg, T.D.H. Arene cis-dihydrodiol formation: From biology to application. Org. Biomol. Chem. 2006, 4, 181–193. [Google Scholar] [CrossRef] [PubMed]
  35. Johnson, R.A. Microbial arene oxidations. In Organic Reactions; John Wiley and Sons, Inc.: New York, NY, USA, 2004; Volume 63, pp. 117–264. [Google Scholar]
  36. Banwell, M.G.; Edwards, A.J.; Harfoot, G.J.; Joliffe, K.A.; McLeod, M.D.; McRae, K.J.; Stewart, S.G.; Vogtle, M. Chemoenzymatic methods for the enantioselective preparation of sesquiterpenoid natural products from aromatic precursors. Pure Appl. Chem. 2003, 75, 223–229. [Google Scholar] [CrossRef]
  37. Boyd, D.R.; Sharma, N.D.; Allen, C.C.R. Aromatic dioxygenases: Molecular biocatalysis and applications. Curr. Opin. Biotechnol. 2001, 12, 564–573. [Google Scholar] [CrossRef]
  38. Froese, J.; Hudlicky, J.R.; Hudlicky, T. Palladium-catalyzed carbonylation of halo arene-cis-dihydrodiols to the corresponding carboxylates. Access to compounds unavailable by toluene dioxygenase-mediated dihydroxylation of the corresponding benzoate esters. Org. Biomol. Chem. 2014, 12, 7810–7819. [Google Scholar] [CrossRef]
  39. Bergmann, M.; Leonidas, Z. Über ein allgemeines verfahren der peptid-synthese. Ber. Dtsch. Chem. Ges. 1932, 65, 1192–1201. [Google Scholar] [CrossRef]
  40. Birkofer, L.; Bierwirth, E.; Ritter, A. Decarbobenzoxylierungen mit triäthylsilan. Chem. Ber. 1961, 94, 821–824. [Google Scholar] [CrossRef]
Figure 1. Naturally occurring Amaryllidaceae constituents (compounds 1 and 2), C-1 derivatives of pancratistatin with enhanced biological activity (compounds 3 and 4) and representative derivatives of narciclasine (compounds 510).
Figure 1. Naturally occurring Amaryllidaceae constituents (compounds 1 and 2), C-1 derivatives of pancratistatin with enhanced biological activity (compounds 3 and 4) and representative derivatives of narciclasine (compounds 510).
Molecules 27 03809 g001
Scheme 1. Synthetic strategy for the synthesis of C-1 narciclasine derivatives.
Scheme 1. Synthetic strategy for the synthesis of C-1 narciclasine derivatives.
Molecules 27 03809 sch001
Scheme 2. Synthesis of the A-ring acid chloride (14) from ortho-vanillin (15). Reagents and conditions: i. Br2, NaOAc, AcOH, rt; ii. H2O2, NaOH, H2O, 0 °C (83% yield from 15); iii. CH2I2, K2CO3, DMF, 80 °C (74% yield); iv. SnCl4, CH3OCHCl2, DCM, −78 °C (66% yield); v. Na2HPO4.H2O, NaClO2, 2-methyl-2-butene, tBuOH:H2O (5:1), rt (92% yield); vi. (COCl)2, DMF, DCM, rt (not purified).
Scheme 2. Synthesis of the A-ring acid chloride (14) from ortho-vanillin (15). Reagents and conditions: i. Br2, NaOAc, AcOH, rt; ii. H2O2, NaOH, H2O, 0 °C (83% yield from 15); iii. CH2I2, K2CO3, DMF, 80 °C (74% yield); iv. SnCl4, CH3OCHCl2, DCM, −78 °C (66% yield); v. Na2HPO4.H2O, NaClO2, 2-methyl-2-butene, tBuOH:H2O (5:1), rt (92% yield); vi. (COCl)2, DMF, DCM, rt (not purified).
Molecules 27 03809 sch002
Scheme 3. Chemoenzymatic oxidation of ortho-dibromobenzene (13) and synthesis of the functionalized oxazine (21). Reagents and conditions: i. toluene dioxygenase in E. coli JM109 (pDTG601A), 35 °C (87%, 2.4 g/L); ii. 2,2-DMP, pTsOH then NaIO4, CbzNHOH, MeOH, H2O, 0 °C to rt (74% yield).
Scheme 3. Chemoenzymatic oxidation of ortho-dibromobenzene (13) and synthesis of the functionalized oxazine (21). Reagents and conditions: i. toluene dioxygenase in E. coli JM109 (pDTG601A), 35 °C (87%, 2.4 g/L); ii. 2,2-DMP, pTsOH then NaIO4, CbzNHOH, MeOH, H2O, 0 °C to rt (74% yield).
Molecules 27 03809 sch003
Scheme 4. Elaboration of the C-ring at C-1 via a Pd-catalyzed carbonylation. Reagents and conditions: i. Al(Hg), THF/H2O (10:1), 0 °C (40% yield); ii. TBSCl, imidazole, DCM, reflux (86% yield); iii. CO, Pd(PPh3)4, Et3N, DMF:MeOH (1:1), 65 °C (54% yield); iv. Et3SiH, PdCl2, Et3N, DCM, rt (not purified).
Scheme 4. Elaboration of the C-ring at C-1 via a Pd-catalyzed carbonylation. Reagents and conditions: i. Al(Hg), THF/H2O (10:1), 0 °C (40% yield); ii. TBSCl, imidazole, DCM, reflux (86% yield); iii. CO, Pd(PPh3)4, Et3N, DMF:MeOH (1:1), 65 °C (54% yield); iv. Et3SiH, PdCl2, Et3N, DCM, rt (not purified).
Molecules 27 03809 sch004
Scheme 5. C-1 functionalization via formation of vinyl nitrile 25 and a proposed mechanism for the formation of compound 26. Reagents and conditions: i. CuCN, DMF, 90 °C (43% yield); ii. DIBAL-H, toluene, −78 °C (45% yield).
Scheme 5. C-1 functionalization via formation of vinyl nitrile 25 and a proposed mechanism for the formation of compound 26. Reagents and conditions: i. CuCN, DMF, 90 °C (43% yield); ii. DIBAL-H, toluene, −78 °C (45% yield).
Molecules 27 03809 sch005
Scheme 6. Intramolecular Heck reaction followed by global deprotection completes the synthesis to form 1-methoxycarbonyl narciclasine (10). Reagents and conditions: i. 14, cat. DMAP, pyridine, DCM, 0 °C to rt (72% yield over two steps); ii. Boc2O, DMAP, CH3CN, rt (75% yield); iii. Pd(OAc)2, dppe, Ag3PO4, toluene, 130 °C, (61% yield brsm); iv. LiCl, DMF, 90 °C (80% yield); v. HCl (1.1 M), THF, 0 °C to rt (55% yield).
Scheme 6. Intramolecular Heck reaction followed by global deprotection completes the synthesis to form 1-methoxycarbonyl narciclasine (10). Reagents and conditions: i. 14, cat. DMAP, pyridine, DCM, 0 °C to rt (72% yield over two steps); ii. Boc2O, DMAP, CH3CN, rt (75% yield); iii. Pd(OAc)2, dppe, Ag3PO4, toluene, 130 °C, (61% yield brsm); iv. LiCl, DMF, 90 °C (80% yield); v. HCl (1.1 M), THF, 0 °C to rt (55% yield).
Molecules 27 03809 sch006
Figure 2. Dose-dependent cytotoxicity of compounds in lung cancer cell line A549. Cells were treated with a series of dilutions of the compounds in 96-well plates for 4 days, and cell viability was measured using MTT assay. Shown in each figure is dose-dependent response to the indicated compound, with the dotted line signifying the IC50.
Figure 2. Dose-dependent cytotoxicity of compounds in lung cancer cell line A549. Cells were treated with a series of dilutions of the compounds in 96-well plates for 4 days, and cell viability was measured using MTT assay. Shown in each figure is dose-dependent response to the indicated compound, with the dotted line signifying the IC50.
Molecules 27 03809 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Habaz, L.; Bedard, K.; Smith, M.; Du, L.; Kornienko, A.; Hudlicky, T. Design and Synthesis of C-1 Methoxycarbonyl Derivative of Narciclasine and Its Biological Activity. Molecules 2022, 27, 3809. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27123809

AMA Style

Habaz L, Bedard K, Smith M, Du L, Kornienko A, Hudlicky T. Design and Synthesis of C-1 Methoxycarbonyl Derivative of Narciclasine and Its Biological Activity. Molecules. 2022; 27(12):3809. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27123809

Chicago/Turabian Style

Habaz, Lihi, Korey Bedard, Mitchell Smith, Liqin Du, Alexander Kornienko, and Tomas Hudlicky. 2022. "Design and Synthesis of C-1 Methoxycarbonyl Derivative of Narciclasine and Its Biological Activity" Molecules 27, no. 12: 3809. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27123809

Article Metrics

Back to TopTop