Next Article in Journal
Dalbergia ecastaphyllum (L.) Taub. and Symphonia globulifera L.f.: The Botanical Sources of Isoflavonoids and Benzophenones in Brazilian Red Propolis
Next Article in Special Issue
Synthetic Strategies, Reactivity and Applications of 1,5-Naphthyridines
Previous Article in Journal
Coumarins as Powerful Photosensitizers for the Cationic Polymerization of Epoxy-Silicones under Near-UV and Visible Light and Applications for 3D Printing Technology
Previous Article in Special Issue
Synthesis and Antibacterial Evaluation of Novel 1,3,4-Oxadiazole Derivatives Containing Sulfonate/Carboxylate Moiety
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Two Annulated Azaheterocyclic Cores Readily Available from a Single Tetrahydroisoquinolonic Castagnoli–Cushman Precursor

Institute of Chemistry, Saint Petersburg State University, 199034 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Submission received: 15 April 2020 / Revised: 23 April 2020 / Accepted: 27 April 2020 / Published: 28 April 2020
(This article belongs to the Collection Heterocyclic Compounds)

Abstract

:
A novel approach to indolo[3,2-c]isoquinoline and dibenzo[c,h][1,6]naphthyridine tetracyclic systems was discovered based on switchable reduction of 2-methoxy-3-(2-nitrophenyl)-1-oxo-1,2,3,4-tetrahydroisoquinoline-4-carboxylic acid prepared via Castagnoli–Cushman reaction. Reduction with ammonium formate resulted in the expected selective transformation of the nitro group (thus providing access to substituted dibenzo[c,h][1,6]naphthyridine via cyclization and dehydrogenation). However, attempted reduction with sodium sulfide initiated a previously unknown reaction cascade including double reduction, cyclization, and decarboxylation, leading to formation of indolo[3,2-c]isoquinoline polyheterocycle in one synthetic step.

Graphical Abstract

1. Introduction

During the last decades, the formal [4 + 2] cycloaddition of homophthalic anhydride (HPA) and imines (recently dubbed the Castagnoli–Cushman reaction, CCR [1,2,3]) has proven itself as a powerful tool for accessing tetrahydroisoquinolonic (THIQ) acids 1 (Figure 1), the promising scaffolds for the design of biologically active compounds. Recently, we reported [4] a practically simple approach to cyclic hydroxamic acids (N-hydroxy THIQ acids 1, R1 = OH) via the CCR, which employed aldoximes as imine surrogates. Hydroxamic acids, in general, are known metal chelators successfully applied in analytical chemistry [5], nuclear fuel processing [6], mineral collection [7], corrosion inhibition [8], and drug design. Metal binding properties often determine the biological activity of hydroxamic acids, which can act as siderophores [9] (bacterial iron transporters) as well as inhibitors of metalloenzymes (such as matrix metalloproteases [10], hyston deacetylases [11], and HIV integrase [12]). In one of our previous works, we demonstrated that conjugates of N-hydroxy THIQ acids (1, R1 = OH) with various fluorophores can serve as selective fluorescent chemosensors for specific metal ions [13].
In this study, we aimed at developing a modified type of such sensors in which attachment of an external fluorescent probe would be unnecessary if the extended π-system of the probe itself would enable it to act as the source of fluorescence. We envisioned that this could be achieved by the annulation of an additional aromatic ring as in, for example, 5-hydroxydibenzo[c,h][1,6]naphthyridine (2) (Figure 2). However, while there are a handful of methods for the preparation of 5-alkyl [14] or 5H- [15,16] dibenzo[c,h][1,6]naphthyridines reported in the literature, the synthesis of the corresponding 5-hydroxy derivatives has not been described. While developing a strategy to construct target compound 2 (the goal which was ultimately achieved), we also arrived, unexpectedly, at a representative of an entirely different heterocyclic system, namely, indolo[3,2-c]isoquinolines 3, which are related to cryptolepine [17], an antimalarial alkaloid from West African plant Cryptolepis sanguinolenta [18,19,20]. Our newly discovered approach to the construction of indolo[3,2-c]isoquinolines is quite distinct from the methods reported in the literature, most of which are based on the transformations of functionalized indoles [21,22] or condensations of two non-annulated starting materials [23,24,25]. Herein, we present the detailed results obtained in the course of this investigation.

2. Results and Discussion

Retrosynthetic analysis of compound 2 with the disconnection of the amide bond brought us to the key intermediate 4 which is based on a typical THIQ scaffold obtainable by the CCR (Figure 3). Since the free amino group would not be compatible with CCR (as it would react with HPA), it was decided to obtain the corresponding nitro derivative 5 first and then reduce it. O-Methyl protecting group on hydroxamate moiety was introduced in order to prevent possible elimination of a water molecule in the course of subsequent elaboration of a lactam ring.
Nitro compound 5 was prepared from HPA and O-methylated oxime 6 (Scheme 1). The reaction was conducted on a gram scale under solvent-free conditions at 110 °C. Such forcing reaction conditions have been previously shown to be crucial for CCR with poorly reactive O-alkyl oximes derived from electron-deficient benzaldehydes [26]. Notably, the yield of CCR product 5 (67%) was substantially higher compared to the yield (36%) obtained in the previously reported reaction of the corresponding unprotected (N-OH) oxime [4]. This increase in yield is likely attributable to the inability of 6 to undergo HPA-mediated dehydration to the respective nitrile, which was the major side reaction for its N-OH counterpart [4]. The trans-configuration of compound 5 was based on comparing the 3JHH values for vicinal methine protons to those reported in the literature [26].
Surprisingly, the classical reduction of 5 with sodium sulfide (Scheme 1, path a) gave a product with no aliphatic protons in the 1H NMR spectrum and only one carbonyl group signal present in the 13C NMR spectrum, which did not correspond to the expected structure 4. The structure of the isolated reaction product, compound 7, which was found to be a known [27] derivative of indolo[3,2-c]isoquinoline, was determined by X-ray diffraction analysis.
The mechanism of transformation 57 can be rationalized as follows. Presumably, the formation of compound 7 begins with reduction of nitro group giving nitroso intermediate A. Under basic conditions (Na2S gives strongly alkaline solutions when dissolved in water), A can undergo decarboxylation with concomitant trapping of the putative carbanion by the nearby nitroso group, resulting in N-hydroxyindoline B. Successive elimination of a water molecule (BC) and tautomerization (CD) could lead to tetracycle D which can, under the reducing conditions, produce compound 7 possessing the contiguous aromatic system (Scheme 2). Noteworthy, such a complex cascade of transformations proceeds with a very high yield of compound 7 (91%).
The proposed mechanism can be supported by the following literature data. Firstly, nitroso compounds have been established as intermediates in the reduction of nitroarenes with inorganic sulfides [28]. Decarboxylation of THIQ acids such as 5 is a known process occurring under similarly forcing conditions [29]. Decarboxylation of postulated intermediate A could be driven forward by the presence of the nearby nitroso functionality which, in turn, intercepts the carbanion generated upon the loss of CO2. Reactions of CH acids with nitroso compounds in the basic medium leading, upon dehydration, to imines have been reported [30,31]. It has also been shown [32] that O-alkyl hydroxamic acids can be reduced to amides by molecular sulfur under basic conditions. Molecular sulfur is always present in reaction mixtures containing Na2S and an oxidant, in this case, the starting material (5).
Several other reducing agents, including sodium dithionite, hydrogen gas+Pd/C, formic acid+Pd/C, ammonium formate+Pd/C, and stannous chloride were tested to perform selective reduction of nitro group in compound 5. All reagents except for ammonium formate, which gave the desired 2-aminophenyl THIQ acid 4 (Scheme 1, path b) in almost quantitative yield (96%), were unsuccessful (Table S1). Noteworthy, both types of reduction were performed on a gram scale, and corresponding products 4 and 7 were obtained in analytically pure form with no need for chromatographic purification.
The desired tetracyclic hydroxamic acid 2 based on the dibenzo[c,h][1,6]naphthyridine core was obtained from THIQ acid 4 in three synthetic steps: lactamization on treatment with propylphosphonic anhydride (T3P) to give compound 8, dehydrogenation of the latter with DDQ and O-demethylation of compound 9 with BBr3. Surprisingly, lactamization 4→8 failed or proved low-yielding with the use of several common amide coupling reagents, including HATU, CDI, DCC, (COCl)2. To the best of our knowledge, compound 2 is the first N-hydroxydibenzo[c,h][1,6]naphthyridine obtained to-date. Compound 7 was elaborated into 1-chloroisoquinoline 10 on treatment with phosphorus oxychloride. Reduction with zinc in acetic acid gave 11H-indolo[3,2-c]isoquinoline 11 which is a known precursor to antiprotozoal compound 12 [33,34] (Scheme 3). Starting from HPA, compound 11 was prepared in an overall yield of 40%, over 4 steps (see Supplementary material). This is in sharp contrast to the previously reported synthetic routes to 11 (8% over 4 steps [35] or 12% over 5 steps [36]). Another advantage of the new route to compound 11 presented herein contrasting it to the previously reported approaches [33,34] is its being free from the use of palladium catalysts, which is an obvious upside from the standpoint of pharmaceutical production [37].

3. Materials and Methods

3.1. General Information

NMR spectra were acquired with a 400 MHz Bruker Avance III spectrometer (400.13 MHz for 1H and 100.61 MHz for 13C) in DMSO-d6 and were referenced to residual solvent proton signals (δH = 2.50) and solvent carbon signals (δC = 39.5). Mass spectra were acquired with a Bruker maXis HRMS-ESI-qTOF spectrometer (electrospray ionization mode, positive ions detection). X-ray single crystal analysis was performed on Agilent Technologies «SuperNova» diffractometer with monochromated Cu Kα radiation. The temperature was kept at 293 K during data collection. Using Olex2 [38], the structure was solved with the SHELXT [39] structure solution program using Intrinsic Phasing and refined with the SHELXL [40] refinement package using Least Squares minimization. TLC was performed on aluminum-backed pre-coated plates (0.25 mm) with silica gel 60 F254 with a suitable solvent system and was visualized using UV fluorescence. Flash column chromatography on silica (Merck, 230–400 mesh) was performed with Biotage Isolera Prime instrument. Preparative HPLC was carried out on Shimadzu LC-20AP instrument, equipped with a spectrophotometric detector, column: Agilent Zorbax prepHT XDB-C18, 5 µm, 21.2 × 150 mm. Oxime 6 was prepared according to the procedure in the literature [41]. Homophthalic anhydride, reagents, and solvents were obtained from commercial sources and were used without further purification. All reactions were performed under air.

3.2. Synthesis

3.2.1. (±)-(3S,4S)-2-Methoxy-3-(2-nitrophenyl)-1-oxo-1,2,3,4-tetrahydroisoquinoline-4-carboxylic acid (5)

A mixture of homophthalic anhydride (2.2 g, 13 mmol) and 2-nitrobenzaldehyde O-methyl oxime 6 (2.4 g, 13 mmol) was thoroughly ground in a mortar and transferred to a screw-cap vial. The reaction mixture was heated at 110 °C for 16 h (oil bath). After cooling to room temperature, the resulting glassy solid was treated with MeOH (20 mL) and sonicated to obtain suspension, which was filtered. The resulting solid was washed with MeOH (10 mL), filtered and dried in air to obtain pure title compound. Yield 3.0 g, 67%, beige solid. 1H NMR (400 MHz, DMSO-d6) δ 13.26 (br.s, 1H), 8.21 (dd, J = 8.1, 1.5 Hz, 1H), 8.01 (dd, J = 7.5, 1.7 Hz, 1H), 7.64 (td, J = 7.6, 1.5 Hz, 1H), 7.60–7.44 (m, 3H), 7.38 (dd, J = 7.3, 1.6 Hz, 1H), 7.00 (dd, J = 7.8, 1.5 Hz, 1H), 6.33 (d, J = 1.3 Hz, 1H), 4.42 (d, J = 1.4 Hz, 1H), 3.80 (s, 3H). 13C NMR (101 MHz, DMSO-d6) δ 171.7, 160.8, 147.9, 134.7, 133.7, 133.1, 132.9, 131.0, 130.1, 128.8, 127.9, 127.9, 127.3, 126.6, 61.8, 59.3, 50.2. HRMS (ESI-TOF) m/z [M + Na]+ calcd for C17H14NaN2O6 365.0744, found 365.0748.

3.2.2. 6,11-Dihydro-5H-indolo[3,2-c]isoquinolin-5-one (7)

To a stirred suspension of compound 5 (1.5 g, 4.4 mmol) in dioxane and water (25 + 25 mL) sodium sulfide nonahydrate (6.3 g, 26 mmol) was added in one portion at room temperature. After heating at 70 °C in an oil bath for 16h the resulting solution was concentrated in vacuo. The residue was suspended in water (50 mL) and filtered. The solid was washed with water (3 × 20mL) and dried in air to provide pure title compound. Yield 1.03 g, 91%, yellow solid. 1H NMR (400 MHz, DMSO-d6) δ 11.85 (br.s, 1H), 8.45 (s, 1H), 8.35 (dd, J = 8.1, 1.3 Hz, 1H), 8.17 (d, J = 7.9 Hz, 1H), 8.04 (d, J = 7.9 Hz, 1H), 7.95–7.75 (m, 1H), 7.61–7.47 (m, 2H), 7.29 (ddd, J = 8.2, 7.0, 1.2 Hz, 1H), 7.18–6.99 (m, 1H). 13C NMR (101 MHz, DMSO-d6) δ 161.3, 137.2, 132.9, 129.7, 129.0, 126.1, 125.0, 124.6, 121.4, 119.4, 119.4, 118.9, 117.2, 112.3. HRMS (ESI-TOF) m/z [M + Na]+ calcd for C15H10NaN2O 257.0685, found 257.0684.
Crystal Data for compound 7 C30H20N4O2 (M = 468.50 g/mol): monoclinic, space group P21/c (no. 14), a = 11.0070(3) Å, b = 15.9054(4) Å, c = 12.3765(4) Å, β = 95.846(3)°, V = 2155.49(11) Å3, Z = 4, T = 293(2) K, μ(CuKα) = 0.744 mm−1, Dcalc = 1.444 g/cm3, 23901 reflections measured (8.074° ≤ 2Θ ≤ 141.34°), 4127 unique (Rint = 0.0518, Rsigma = 0.0297) which were used in all calculations. The final R1 was 0.0686 (I > 2σ(I)) and wR2 was 0.1995 (all data).

3.2.3. (±)-(3S,4S)-2-Methoxy-3-(2-aminophenyl)-1-oxo-1,2,3,4-tetrahydroisoquinoline-4-carboxylic acid (4)

A round-bottom flask was charged with compound 5 (1.4 g, 4 mmol), 10% wt. Pd/C (424 mg, 0.4 mmol), ammonium formate (2.6 g, 41 mmol) and MeOH (150 mL). The resulting mixture was stirred with reflux for 16 h. After cooling to room temperature, it was filtered through a pad of Celite and filtrate was concentrated in vacuo to provide pure title compound. Yield 1.2 g, 96%, beige solid. 1H NMR (400 MHz, DMSO-d6) δ 8.06–7.83 (m, 1H), 7.42–7.33 (m, 1H), 7.34–7.24 (m, 1H), 7.10 (d, J = 7.5 Hz, 1H), 7.00–6.78 (m, 1H), 6.69 (d, J = 7.9 Hz, 1H), 6.51 (d, J = 7.6 Hz, 1H), 6.37 (t, J = 7.4 Hz, 1H), 5.78 (s, 1H), 5.13 (br.s, 2H), 3.72 (s, 3H), 3.18 (s, 1H). 13C NMR (101 MHz, DMSO-d6, 353 K) δ 174.1, 161.3, 145.2, 137.4, 131.6, 130.8, 128.3, 127.9, 126.6, 125.1, 124.0, 116.5, 116.1, 74.6, 61.5, 61.2, 52.9. HRMS (ESI-TOF) m/z [M + Na]+ calcd for C17H16NaN2O4 335.1002, found 335.0972.

3.2.4. 5-Methoxy-4b,12-dihydrodibenzo[c,h][1,6]naphthyridine-6,11(5H,10bH)-dione (8)

To a stirred suspension of compound 4 (100 mg, 0.32 mmol) in dry chloroform (10 mL) a 50% wt. solution of propanephosphonic acid anhydride, T3P (410 mg, 0.64 mmol) in toluene was added at room temperature. After stirring for 16 h the reaction mixture was diluted with chloroform and st. NaHCO3. The organic layer was separated, washed with water, dried and concentrated to give crude product (85 mg, 90%), which was further purified using column chromatography on silica (eluent CHCl3-MeOH, 0–100% of MeOH). Yield 60 mg, 63%, white solid. 1H NMR (400 MHz, DMSO-d6) δ 10.29 (s, 1H), 8.14 (dd, J = 7.9, 1.5 Hz, 1H), 8.10 (dd, J = 8.0, 1.3 Hz, 1H), 7.83 (d, J = 7.9 Hz, 1H), 7.53 (td, J = 7.6, 1.5 Hz, 1H), 7.43–7.33 (m, 1H), 7.20 (t, J = 7.5 Hz, 1H), 7.04 (td, J = 7.7, 1.3 Hz, 1H), 7.00 (dd, J = 7.9, 1.3 Hz, 1H), 5.16 (d, J = 14.7 Hz, 1H), 4.03 (d, J = 14.7 Hz, 1H), 3.90 (s, 3H). 13C NMR (101 MHz, DMSO-d6) δ 167.8, 138.0, 134.9, 132.9, 130.2, 129.4, 128.7, 128.4, 127.8, 125.8, 122.8, 115.7, 79.6, 63.0, 59.7, 45.8 (signal of one carbonyl group is missing and several signals are broadened due to conformational exchange). HRMS (ESI-TOF) m/z [M + Na]+ calcd for C17H14NaN2O3 317.0897, found 317.0912.

3.2.5. 5-Methoxydibenzo[c,h][1,6]naphthyridine-6,11(5H,12H)-dione (9)

A screw-cap vial was charged with compound 8 (150 mg, 0.51 mmol), 2,3-dichloro-5,6-dicyano-1,4-benzoquinone, DDQ (580 mg, 2.55 mmol) and dry 1,4-dioxane (3 mL). The resulting mixture was heated at 110 °C (oil bath) for 48 h (reaction progress was monitored by TLC). After cooling to room temperature, the reaction mixture was partitioned between ethyl acetate (150 mL) and water, containing ascorbic acid (50 mL + 1 g). The organic layer was separated, washed with water, dried and concentrated to provide pure title compound. Yield 114 mg, 77%, beige solid. 1H NMR (400 MHz, DMSO-d6) δ 12.11 (s, 1H), 9.91 (d, J = 8.5 Hz, 1H), 8.90 (d, J = 8.6 Hz, 1H), 8.53–8.17 (m, 1H), 7.89 (ddd, J = 8.7, 7.1, 1.6 Hz, 1H), 7.74–7.56 (m, 2H), 7.49 (dd, J = 8.3, 1.4 Hz, 1H), 7.31 (ddd, J = 8.5, 7.0, 1.4 Hz, 1H). 13C NMR (101 MHz, DMSO-d6) δ 160.5, 158.6, 141.9, 139.0, 133.9, 133.5, 131.9, 128.3, 127.7, 127.7, 127.3, 125.7, 122.7, 116.5, 111.3, 105.5, 63.7. HRMS (ESI-TOF) m/z [M + Na]+ calcd for C17H12NaN2O3 315.0740, found 315.0740.

3.2.6. 5-Hydroxydibenzo[c,h][1,6]naphthyridine-6,11(5H,12H)-dione (2)

A suspension of compound 9 (20 mg, 0.07 mmol) in dry DCM (4 mL) was cooled to 0 °C in an ice bath followed by addition of solution of BBr3 (102 mg, 0.41 mmol) in dry DCM (1 mL). The mixture was stirred at 0 °C for 3 h and was then allowed to warm to room temperature. The reaction was quenched with sat. NaHCO3 (5 mL) under cooling with ice. The precipitate formed was filtered, washed with water and dried to give crude product (20 mg), which was purified via preparative reverse-phase HPLC eluting with water–acetonitrile (both containing 0.1% TFA). Method: 10 min of 20% ACN, then linear gradient 20–95% ACN for 35 min, flow rate—12 mL/min, column temperature—40 °C. Yield 12 mg, 60%, white glassy solid. 1H NMR (400 MHz, DMSO-d6) δ 12.08 (s, 1H), 12.03 (s, 1H), 9.94 (dd, J = 8.6, 1.1 Hz, 1H), 9.18 (dd, J = 8.7, 1.3 Hz, 1H), 8.41 (dd, J = 7.9, 1.5 Hz, 1H), 7.89 (ddd, J = 8.6, 7.1, 1.6 Hz, 2H), 7.70 (ddd, J = 8.1, 7.1, 1.1 Hz, 1H), 7.63 (ddd, J = 8.3, 7.0, 1.2 Hz, 1H), 7.46 (dd, J = 8.3, 1.2 Hz, 1H), 7.26 (ddd, J = 8.5, 7.0, 1.3 Hz, 1H). 13C NMR (101 MHz, DMSO-d6) δ 160.7, 159.4, 141.9, 138.9, 133.5, 133.3, 131.9, 129.2, 128.2, 127.6, 127.0, 124.3, 121.8, 116.2, 111.9, 104.1. HRMS (ESI-TOF) m/z [M + Na]+ calculated for C16H10NaN2O3 301.0584, found 301.0584.

3.2.7. 5-Chloro-11H-indolo[3,2-c]isoquinoline (10)

Compound 7 (180 mg, 0.77 mmol) was heated with stirring in POCl3 (3 mL) at 100 °C (oil bath) for 16 h in a screw-cap vial. Conversion of starting material was controlled by TLC. After cooling to room temperature, the reaction mixture was carefully poured into ice (25 mL) with vigorous stirring. The resulting solid was filtered, washed with water until neutral pH of washings and dried in air to provide pure title compound. Yield 165 mg, 85%, yellow solid. 1H NMR (400 MHz, DMSO-d6) δ 12.63 (s, 1H), 8.62 (d, J = 8.2 Hz, 1H), 8.42 (d, J = 8.5 Hz, 1H), 8.19 (d, J = 7.9 Hz, 1H), 8.02 (t, J = 7.5 Hz, 1H), 7.82 (t, J = 7.7 Hz, 1H), 7.72 (d, J = 8.2 Hz, 1H), 7.52 (t, J = 7.6 Hz, 1H), 7.33 (t, J = 7.5 Hz, 1H). 13C NMR (101 MHz, DMSO-d6) δ 142.0, 139.1, 132.1, 131.6, 128.1, 127.9, 127.4, 126.5, 126.0, 124.0, 122.6, 122.1, 120.8, 119.7, 112.6. HRMS (ESI-TOF) m/z [M + H]+ calcd for C15H10ClN2 253.0527, found 253.0528.

3.2.8. 11H-Indolo[3,2-c]isoquinoline (11)

A solution of compound 10 (200 mg, 0.79 mmol) in AcOH-water (20 + 2 mL) was heated to 75 °C followed by addition of Zn dust (464 mg, 7.14 mmol) in three portions during 1.5 h (conversion of starting material was monitored by TLC). After cooling to room temperature acid was neutralized with 10% aq. NaOH followed by extraction with DCM (3 × 50 mL). Combined organic layer was dried over Na2SO4, filtered and concentrated to give crude product, which was purified using column chromatography on silica eluting with DCM-MeOH (20:1). Yield 134 mg, 77%, light yellow solid. 1H NMR (400 MHz, DMSO-d6) δ 11.71 (s, 1H), 9.02 (s, 1H), 8.42 (dd, J = 8.3, 1.1 Hz, 1H), 8.26 (d, J = 7.8 Hz, 1H), 8.09 (d, J = 8.2 Hz, 1H), 7.83–7.68 (m, 1H), 7.65–7.50 (m, 3H), 7.44–7.34 (m, 1H), 7.29–7.18 (m, 1H). 13C NMR (101 MHz, DMSO-d6) δ 145.1, 139.0, 134.0, 130.4, 129.1, 127.8, 127.0, 126.6, 126.0, 123.9, 123.2, 121.6, 120.3, 119.7, 112.3. HRMS (ESI-TOF) m/z [M + H]+ calcd for C15H11N2 219.0917, found 219.0908.

4. Conclusions

In summary, we have expanded the scope of heterocyclic scaffolds accessible though the Castagnoli–Cushman reaction by developing protocols for the switchable reduction of 2-methoxy-3-(2-nitrophenyl)-1-oxo-1,2,3,4-tetrahydroisoquinoline-4-carboxylic acid, leading either to novel tetracyclic hydroxamic acid with dibenzo[c,h][1,6]naphthyridine core (following lactamization and deprotection) or to indolo[3,2-c]isoquinolines (as the result of hitherto unknown cascade heteroannulation reaction). Metal binding properties of dibenzo[c,h][1,6]naphthyridine-based hydroxamic acid obtained and its ability to mimic bacterial siderophores are currently under investigation in our laboratories. The route to antiprotozoal indolo[3,2-c]isoquinolines discovered in the course of this study is the most high-yielding among Pd-free protocols reported in the literature.

Supplementary Materials

The following are available online, copies of 1H and 13C spectra for all prepared compounds, results of reagent screening for reduction of compound 5 (Table S1) and crystallographic data for compound 7 (Figure S1, Table S2). CCDC 1996076 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via http://www.ccdc.cam.ac.uk.

Author Contributions

Organic synthesis and compound characterization, E.K. and O.B.; conceptualization and project administration, O.B. and D.D.; methodology and validation, O.B. and D.D.; Writing—Original Draft preparation, O.B. and M.K.; Writing—Review and Editing, M.K. and D.D.; funding acquisition, O.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Russian Science Foundation project grant no. 18-73-00074.

Acknowledgments

The authors are grateful to the Research Centre for Magnetic Resonance, the Centre for Chemical Analysis and Materials Research, and the Centre for X-ray Diffraction Methods of Saint Petersburg State University Research Park for the analytical data.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Gonzalez-Lopez, M.; Shaw, J.T. Cyclic anhydrides in formal cycloadditions and multicomponent reactions. Chem. Rev. 2009, 109, 164–189. [Google Scholar] [CrossRef] [PubMed]
  2. Krasavin, M.; Dar’in, D. Current diversity of cyclic anhydrides for the Castagnoli–Cushman-type formal cycloaddition reactions: Prospects and challenges. Tetrahedron Lett. 2016, 57, 1635–1640. [Google Scholar] [CrossRef]
  3. Cushman, M.; Dikshit, D.K. Formation of the 5-benzo[d]naphtho[2,3-b]pyran system during an attempted benzophenanthridine synthesis. J. Org. Chem. 1980, 45, 5064–5067. [Google Scholar] [CrossRef]
  4. Bakulina, O.; Bannykh, A.; Dar’in, D.; Krasavin, M. Cyclic Hydroxamic Acid Analogues of Bacterial Siderophores as Iron-Complexing Agents prepared through the Castagnoli-Cushman Reaction of Unprotected Oximes. Chem. Eur. J. 2017, 23, 17667–17673. [Google Scholar] [CrossRef]
  5. Abdein, M.A.; Abualreish, M.J.A. The Analytical Applications and Biological Activity of Hydroxamic acids. J. Adv. Chem. 2014, 10, 2118–2125. [Google Scholar]
  6. Birkett, J.E.; Carrott, M.J.; Fox, O.D.; Jones, C.J.; Maher, C.J.; Roube, C.V.; Taylor, R.J.; Woodhead, D.A. Controlling Neptunium and Plutonium within Single Cycle Solvent Extraction Flowsheets for Advanced Fuel Cycles. J. Nucl. Sci. Technol. 2007, 44, 337–343. [Google Scholar] [CrossRef]
  7. Natarajan, R. Hydroxamic Acids as Chelating Mineral Collectors. In Hydroxamic Acids; Gupta, S.P., Ed.; Springer Berlin Heidelberg: Berlin/Heidelberg, Germany, 2013; pp. 281–307. [Google Scholar] [CrossRef]
  8. Ezznaydy, G.; Shaban, A.; Telegdi, J.; Ouaki, B.; El Hajjaji, S. Inhibition of copper corrosion in saline solution by mono-hydroxamic acid. J. Mater. Environ. Sci. 2015, 6, 1819–1823. [Google Scholar]
  9. Schalk, I.J.; Hannauer, M.; Braud, A. New roles for bacterial siderophores in metal transport and tolerance. Environ. Microbiol. 2011, 13, 2844–2854. [Google Scholar] [CrossRef]
  10. Kapranov, L.E.; Reznikov, A.N.; Klimochkin, Y.N. The Main Structural Types of Inhibitors of Matrix Metalloproteinases. Pharm. Chem. J. 2017, 51, 175–181. [Google Scholar] [CrossRef]
  11. Manal, M.; Chandrasekar, M.J.; Gomathi Priya, J.; Nanjan, M.J. Inhibitors of histone deacetylase as antitumor agents: A critical review. Bioorg. Chem. 2016, 67, 18–42. [Google Scholar] [CrossRef]
  12. Plewe, M.B.; Butler, S.L.; Dress, K.R.; Hu, Q.; Johnson, T.W.; Kuehler, J.E.; Kuki, A.; Lam, H.; Liu, W.; Nowlin, D.; et al. Azaindole hydroxamic acids are potent HIV-1 integrase inhibitors. J. Med. Chem. 2009, 52, 7211–7219. [Google Scholar] [CrossRef] [PubMed]
  13. Bakulina, O.; Rashevskii, A.; Dar’in, D.; Halder, S.; Khagar, P.; Krasavin, M. Modular Assembly of Tunable Fluorescent Chemosensors Selective for Pb2+ and Cu2+ Metal Ions via the Multicomponent Castagnoli-Cushman Reaction. ChemistrySelect 2019, 4, 6066–6073. [Google Scholar] [CrossRef]
  14. Kiselev, E.; Dexheimer, T.S.; Pommier, Y.; Cushman, M. Design, synthesis, and evaluation of dibenzo[c,h][1,6]naphthyridines as topoisomerase I inhibitors and potential anticancer agents. J. Med. Chem. 2010, 53, 8716–8726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Mrkvicka, V.; Klasek, A.; Kimmel, R.; Pevee, A.; Kosmrlj, J. Thermal reaction of 3aH,5H-thiazolo[5,4-c]quinoline-2,4-diones–an easy pathway to 4-amino-1H-quinolin-2-ones and novel 6H-thiazolo[3,4-c]quinazoline-3,5-diones. Arkivoc 2008, 2008, 289–302. [Google Scholar]
  16. Stadlbauer, W.; Kappe, T. Synthese von Indolen und Isochinolonen aus Phenylmalonylheterocyclen. Monatsh. Chem. 1984, 115, 467–475. [Google Scholar] [CrossRef]
  17. Delvaux, E. Cryptolepine. J. Pharm. Belg. 1931, 13, 955. [Google Scholar]
  18. Lavrado, J.; Moreira, R.; Paulo, A. Indoloquinolines as scaffolds for drug discovery. Curr. Med. Chem. 2010, 17, 2348–2370. [Google Scholar] [CrossRef]
  19. Parvatkar, P.T.; Parameswaran, P.S.; Tilve, S.G. Isolation, biological activities, and synthesis of indoloquinoline alkaloids: Cryptolepine, isocryptolepine, and neocryptolepine. Curr. Org. Chem. 2011, 15, 1036–1057. [Google Scholar] [CrossRef] [Green Version]
  20. Wright, C.W. Recent developments in naturally derived antimalarials: Cryptolepine analogues. J. Pharm. Pharmacol. 2007, 59, 899–904. [Google Scholar] [CrossRef]
  21. Fodor, L.; Csomós, P.; Csámpai, A.; Sohár, P. Novel indole syntheses by ring transformation of β-lactam-condensed 1,3-benzothiazines into indolo[2,3-b][1,4]benzothiazepines and indolo[3,2-c]isoquinolines. Tetrahedron 2012, 68, 851–856. [Google Scholar] [CrossRef]
  22. Hu, Z.; Tong, X.; Liu, G. Rhodium(III)-Catalyzed Cascade Cyclization/Electrophilic Amidation for the Synthesis of 3-Amidoindoles and 3-Amidofurans. Org. Lett. 2016, 18, 2058–2061. [Google Scholar] [CrossRef]
  23. Nguyen, H.H.; Fettinger, J.C.; Haddadin, M.J.; Kurth, M.J. Expedient one-pot synthesis of indolo[3,2-c]isoquinolines via a base-promoted N-alkylation/tandem cyclization. Tetrahedron Lett. 2015, 56, 5429–5433. [Google Scholar] [CrossRef] [Green Version]
  24. Jagtap, P.G.; Baloglu, E.; Southan, G.; Williams, W.; Roy, A.; Nivorozhkin, A.; Landrau, N.; Desisto, K.; Salzman, A.L.; Szabo, C. Facile and convenient syntheses of 6,11-dihydro-5H-indeno[1,2-c]isoquinolin- 5-ones and 6,11-dihydro-5H-indolo[3,2-c]isoquinolin-5-one. Org. Lett. 2005, 7, 1753–1756. [Google Scholar] [CrossRef]
  25. Li, L.; Chua, W.K.S. One-pot multistep synthesis of 3,4-fused isoquinolin-1(2H)-one analogs. Tetrahedron Lett. 2011, 52, 1574–1577. [Google Scholar] [CrossRef]
  26. Chupakhin, E.; Bakulina, O.; Dar’in, D.; Krasavin, M. Facile Access to Fe(III)-Complexing Cyclic Hydroxamic Acids in a Three-Component Format. Molecules 2019, 24, 864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Hiremath, S.P.; Saundane, A.R.; Mruthyunjayaswamy, B.H.M. A new method for the synthesis of 6H,11H-indolo[3,2-c]-isoquinolin-5-ones/thiones and their reactions. J. Heterocycl. Chem. 1993, 30, 603–609. [Google Scholar] [CrossRef]
  28. Cope, O.J.; Brown, R.K. The Reduction of Nitrobenzene by Sodium Sulphide in Aqueous Ethanol. Can. J. Chem. 1961, 39, 1695–1710. [Google Scholar] [CrossRef]
  29. Kiselev, E.; Empey, N.; Agama, K.; Pommier, Y.; Cushman, M. Dibenzo[c,h][1,5]naphthyridinediones as topoisomerase I inhibitors: Design, synthesis, and biological evaluation. J. Org. Chem. 2012, 77, 5167–5172. [Google Scholar] [CrossRef] [Green Version]
  30. Nohira, H.; Sato, K.; Mukaiyama, T. The Reactions of Nitrosobenzene and Some Methylene Compounds. B. Chem. Soc. Jpn. 1963, 36, 870–872. [Google Scholar] [CrossRef]
  31. Payette, J.N.; Yamamoto, H. Nitrosobenzene-mediated C-C bond cleavage reactions and spectral observation of an oxazetidin-4-one ring system. J. Am. Chem. Soc. 2008, 130, 12276–12278. [Google Scholar] [CrossRef] [Green Version]
  32. Wang, S.; Zhao, X.; Zhang-Negrerie, D.; Du, Y. Reductive cleavage of the N–O bond: Elemental sulfur-mediated conversion of N-alkoxyamides to amides. Org. Chem. Front. 2019, 6, 347–351. [Google Scholar] [CrossRef]
  33. Van Baelen, G.; Hostyn, S.; Dhooghe, L.; Tapolcsanyi, P.; Matyus, P.; Lemiere, G.; Dommisse, R.; Kaiser, M.; Brun, R.; Cos, P.; et al. Structure-activity relationship of antiparasitic and cytotoxic indoloquinoline alkaloids, and their tricyclic and bicyclic analogues. Bioorg. Med. Chem. 2009, 17, 7209–7217. [Google Scholar] [CrossRef]
  34. Van Baelen, G.; Meyers, C.; Lemière, G.L.F.; Hostyn, S.; Dommisse, R.; Maes, L.; Augustyns, K.; Haemers, A.; Pieters, L.; Maes, B.U.W. Synthesis of 6-methyl-6H-indolo[3,2-c]isoquinoline and 6-methyl-6H-indolo[2,3-c]isoquinoline: Two new unnatural isoquinoline isomers of the cryptolepine series. Tetrahedron 2008, 64, 11802–11809. [Google Scholar] [CrossRef]
  35. Martin, M.J.; Trudell, M.L.; Diaz Arauzo, H.; Allen, M.S.; LaLoggia, A.J.; Deng, L.; Schultz, C.A.; Tan, Y.C.; Bi, Y.; Narayanan, K.; et al. Molecular yardsticks. Rigid probes to define the spatial dimensions of the benzodiazepine receptor binding site. J. Med. Chem. 1992, 35, 4105–4117. [Google Scholar] [CrossRef] [PubMed]
  36. Béres, M.; Timári, G.; Hajós, G. Straightforward synthesis of 11H-indolo[3,2-c]isoquinoline and benzofuro[3,2-c]isoquinoline by ring transformation. Tetrahedron Lett. 2002, 43, 6035–6038. [Google Scholar] [CrossRef]
  37. Hayler, J.D.; Leahy, D.K.; Simmons, E.M. A Pharmaceutical Industry Perspective on Sustainable Metal Catalysis. Organometallics 2018, 38, 36–46. [Google Scholar] [CrossRef]
  38. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  39. Sheldrick, G.M. SHELXT-integrated space-group and crystal-structure determination. Acta Crystallogr. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  40. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  41. Salerno, C.P.; Resat, M.; Magde, D.; Kraut, J. Synthesis of Caged NAD(P)+Coenzymes: Photorelease of NADP+. J. Am. Chem. Soc. 1997, 119, 3403–3404. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds 4, 5 and 7 are available from the authors.
Figure 1. Synthesis of tetrahydroisoquinolonic (THIQ) acids via the Castagnoli–Cushman reaction and its mechanism.
Figure 1. Synthesis of tetrahydroisoquinolonic (THIQ) acids via the Castagnoli–Cushman reaction and its mechanism.
Molecules 25 02049 g001
Figure 2. THIQ acids 1 and related scaffolds.
Figure 2. THIQ acids 1 and related scaffolds.
Molecules 25 02049 g002
Figure 3. Retrosynthetic analysis N-hydroxy dibenzo[c,h][1,6]naphthyridine 2.
Figure 3. Retrosynthetic analysis N-hydroxy dibenzo[c,h][1,6]naphthyridine 2.
Molecules 25 02049 g003
Scheme 1. Switchable reduction of CCR product 5 and crystal structure of unexpected product 7 (CCDC 1996076).
Scheme 1. Switchable reduction of CCR product 5 and crystal structure of unexpected product 7 (CCDC 1996076).
Molecules 25 02049 sch001
Scheme 2. Plausible reaction mechanism of transformation 57.
Scheme 2. Plausible reaction mechanism of transformation 57.
Molecules 25 02049 sch002
Scheme 3. Post-condensational modifications of compounds (a) 4 and (b) 7.
Scheme 3. Post-condensational modifications of compounds (a) 4 and (b) 7.
Molecules 25 02049 sch003

Share and Cite

MDPI and ACS Style

Karchuganova, E.; Bakulina, O.; Dar’in, D.; Krasavin, M. Two Annulated Azaheterocyclic Cores Readily Available from a Single Tetrahydroisoquinolonic Castagnoli–Cushman Precursor. Molecules 2020, 25, 2049. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25092049

AMA Style

Karchuganova E, Bakulina O, Dar’in D, Krasavin M. Two Annulated Azaheterocyclic Cores Readily Available from a Single Tetrahydroisoquinolonic Castagnoli–Cushman Precursor. Molecules. 2020; 25(9):2049. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25092049

Chicago/Turabian Style

Karchuganova, Elizaveta, Olga Bakulina, Dmitry Dar’in, and Mikhail Krasavin. 2020. "Two Annulated Azaheterocyclic Cores Readily Available from a Single Tetrahydroisoquinolonic Castagnoli–Cushman Precursor" Molecules 25, no. 9: 2049. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25092049

Article Metrics

Back to TopTop