Next Article in Journal
Evolution of the Electronic Structure of the trans-[Re6S8bipy4Cl2] Octahedral Rhenium Cluster during Reduction
Previous Article in Journal
Phenolic, Headspace and Sensory Profile, and Antioxidant Capacity of Fruit Juice Enriched with Salvia officinalis L. and Thymus serpyllum L. Extract: A Potential for a Novel Herbal-Based Functional Beverages
Previous Article in Special Issue
Synthesis and Cytotoxic Activity of the Derivatives of N-(Purin-6-yl)aminopolymethylene Carboxylic Acids and Related Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Two-Step Synthesis of Unprotected 3-Aminoindoles via Post Functionalization with Nitrostyrene

by
Nicolai A. Aksenov
*,
Nikolai A. Arutiunov
,
Igor A. Kurenkov
,
Vladimir V. Malyuga
,
Dmitrii A. Aksenov
,
Daria S. Momotova
,
Anna M. Zatsepilina
,
Elizaveta A. Chukanova
,
Alexander V. Leontiev
and
Alexander V. Aksenov
*
Department of Chemistry, North Caucasus Federal University, 1a Pushkin St., 355009 Stavropol, Russia
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(9), 3657; https://doi.org/10.3390/molecules28093657
Submission received: 30 March 2023 / Revised: 19 April 2023 / Accepted: 21 April 2023 / Published: 23 April 2023
(This article belongs to the Special Issue Synthesis of Bioactive Compounds)

Abstract

:
A novel, low-cost method for the preparation of not easily accessible free 3-aminoindoles has been developed. This approach is based on a well-established reaction between indoles and nitrostyrene in the presence of phosphorous acid, which results in the formation of 4′-phenyl-4′H-spiro[indole-3,5′-isoxazoles]. The latter could be transformed to corresponding aminated indoles by reaction with hydrazine hydrate in good or excellent yields upon microwave-assisted heating.

1. Introduction

3-aminoindolic motif commonly occurs in many natural and artificial substances demonstrating a wide range of biological activities [1,2,3,4,5]. Ideally, to install this fragment into the target molecules, one would need a convenient, broadly applicable synthetic pathway to free, 3-aminated indoles as corresponding precursors or building blocks. However, the instability of the unprotected, electron-rich 3-aminoindoles, which are sensitive to light and air and tend to undergo oxidative dimerization and/or other types of decomposition reactions [6,7,8] is the main reason why such a method is still yet to be found. Until then, most of the reported synthesizes toward 3-aminoindoles are usually dealt with relatively stable electron-pour-deactivated derivatives [9,10,11,12] or rely on capping the in situ generated amino group with suitable protective groups [13,14,15,16,17]. Up to date, there are just a few published procedures where unprotected 3-aminoindoles were isolated and characterized [7,8,18].
On the other hand, common approaches to 3-aminoindole derivatives can be roughly divided into two kinds: non-indolic methods, which build up the aminoindole skeleton from scratch via the Fisher [19,20] or similar multicomponent annulation reactions [7,8,9,10,11,12,14,21,22,23,24] and post-functionalization procedures based on the corresponding 3-substituted indoles. As for the latter, the major strategies for introducing an amino group at the C3 indole position have remained the nitration [2,13,15,17,25,26] or azidation [18] reactions followed by a reduction to the free amine. The synthesis starts from the corresponding 3-indolecarboxylic acids by a two-step sequence involving the Curtius rearrangement [16,27], palladium-catalyzed amination of indole halides [28], as well as a number of some recent direct C-H amination methods [29,30,31,32,33,34], have been also reported. However, these transformations are mostly multistep processes that require subsequent protection–deprotection or functional group interconversion steps and often suffer from limited scope and efficacy. In turn, herein, we would like to present a novel two-step method for the preparation of unprotected 2-aryl-3-aminoindoles 5 directly from the corresponding 2-aryl indoles 1 and nitrostyrene 2 via intermedial spirocyclic isoxazoles 3 or indolinone 4 (Scheme 1). This approach provides a straightforward synthetic route to otherwise not easily accessible free 3-aminoindoles.

2. Results

A few years ago, we have discovered [35,36] a somewhat unusual reaction between indoles 6 and nitrostyrenes 7, where the latter act as 1,4-dipoles in the presence of phosphorous acid to give diastereomerically pure spiro-2-indolinone isoxazoles 8 in good to excellent yields (Scheme 2). Subsequent treatment with a mild acid or base leads to 2-(3-oxoindolin-2-yl)-2-arylacetonitriles 9 [37,38], which are due to the presence of versatile cyano and carbonyl functional groups could serve as a good synthetic platform for carrying out many other useful transformations.
Thus, so far, we have shown that upon the action of KOH in refluxing ethanol, the N-alkyl indolinones 9 formed pyrroloindoles 10, while NH derivatives under the same conditions demonstrated the unexpected extrusion of arylacetonitrile molecule, which eventually result in the formation of hydroxyindolinones 11 [39]. The reduction of 8 or 9 with sodium borohydride has proved to be an efficient way of preparation of corresponding indolylacetamides 12 [40], and a reaction of 2-(3-oxoindolin-2-yl)-2-arylacetonitriles 9 with benzene-1,2-diamines 13 furnished quinoxalines 14 in high-yields [41].
It should be noticed that the latter transformation is accompanied by loss of phenylacetonitrile molecule, so, at some point, we speculated about using hydrazine hydrate instead of o-phenylenediamine derivatives to avoid it. In this case, starting, for instance, from indolinone 4aa, one would expect aminopyridazine 15 as a product analogously to the results of the work [42] (Scheme 3). The authors of that work used the structurally similar substrate 16 to obtain pyridazinoindole 17 by cyclocondensation, the former with hydrazine hydrate in boiling acetic acid.
Keeping that in mind, firstly, we tried to reflux 4aa in hydrazine hydrate, only to get the starting material back unchanged. In the attempt to force the reaction, microwave-assisted heating at 200 °C for 1 h was applied, and this time, the conversion of 4aa did occur, affording, to our surprise, 3-amino-2-phenylindole 5aa as the only isolable product (Scheme 4). In turn, the N-methylated indolinone 18 under the same conditions furnished the corresponding N-Me 3-aminoindole 19, although in a rather modest 41% yield.
Apparently, as opposed to our initial hypothesis, the use of hydrazine hydrate (Entry 9) with 4aa does not lead to the target aminopyridazine 15 nor other hydrazines (Entries 3–8), and nitrogen binucleophiles (Entries 1,2,4–9) lead each time to the 3-aminoindole 5aa (Table 1).
To come up with a plausible mechanism, one had to accommodate all those observations above, namely the formation of the 3-aminoindole 5aa with both R3CH2NH2 (Entries 1,2) and hydrazine (Entries 3–9) derivatives while lacking any expected products in the case of hydroxylamine (Entry 10). We speculate (Scheme 5) that for the amine-bearing α-CH2 group (such as ethylene or 1,2-propylenediamines), the condensation of the latter with the starting indolinone A leads to the corresponding imine B. The following proton abstraction from the N-alkylimino fragment, accompanied by simultaneous benzyl cyanide loss, gives a new, aromatization-driven imine C. Then, due to an excess of R3CH2NH2, a recondensation occurs, resulting in the target aminated indole D. In favor of this mechanism, a key feature of which is the presence of α-CH protons, speaks to our previous finding [41], where a reaction of (3-oxoindolin-2-yl)acetonitriles A with 1,2-phenylenediamines ends up in quinoxalines 14 (Scheme 2). If anything like that were to happen in the case of 1,2-diamines (Entries 1,2), then the formation of the corresponding dihydroquinaxolines was expected but did not occur.
Meanwhile, in the case of hydrazine derivatives (Entries 3–9), supposedly, the reaction takes a slightly different pathway (Scheme 6). First of all, like with imine B (Scheme 5), the formation of hydrazone E should occur. The following extrusion of a phenylacetonitrile molecule would lead to the azo-heteroarene F. As shown previously, the -N=N- bond undergoes a reductive cleavage to the corresponding anilines upon the action of (among other reductants) hydrazine hydrate in the presence of, for instance, Raney Ni [43], aluminum powder [44], or even without any catalyst by simple heating in ethanol [45]. We assume that something similar takes place in our case resulting eventually in the target 3-aminoindole D.
Lastly, in the reaction with hydroxylamine (Entry 10), no meaningful products were isolated. However, the N-methyl derivative 18, under the same conditions, unexpectedly gave 1-methyl-2-phenylindole J, although with a low 15% yield (Scheme 7). Most likely, the corresponding oxime G, after the loss of the BnCN molecule, became nitrosoindole H. Further hydrolysis of benzyl cyanide and its condensation with the nitroso group of indole H produced the azo compound I, which in turn got reduced in some Wolff–Kishner-type reaction to form the 1-methyl-2-phenylindole J.
Next, we evaluated the scope and limitations of the described procedure (Entry 9). For that, a series of indolinones 4 bearing various aryl substituents R1 was introduced into the reaction with hydrazine hydrate under the chosen conditions (Method A). As seen in Scheme 8, all these substrates reacted smoothly, producing the corresponding products 5aa5ag in good to high yields. The presence of alkyl or fluorine substitutes at C-5 in the indoline core did not affect the reaction performance, and the target 5-substituted 3-aminoindoles 5ab, 5ac, and 5ad were also obtained in high yields (Scheme 8). Remarkably, the direct conversion of spiranes 3 into aminoindoles 5 (Method B) is also possible, giving yields comparable to those obtained via Method A.
Finally, we tested the possibility of subsequent protection of the newly formed amino group by running a reaction in the mixture of hydrazine hydrate and acetic acid. Expectedly, the corresponding N-acyl derivatives N-Ac 5aa, 5ai were received, although in somewhat reduced yields (Scheme 9).
In the end, we would like to share some of our thoughts and observations regarding the shelf stability of the described herein free 3-aminoindoles. It seems that there is a general consensus about the sensitivity of these substances to air and light both in solution and solid and, as a result, the inability to purify them by column chromatography [7,8]. At the same time, we were able to work with most of them (but not sample 19) rather comfortably, including short and fast (10–15 min) column purifications on silica. The freshly prepared samples, usually light grey or beige, indeed became deep blue with time but even then, their proton NMR spectra showed no signs of significant decomposition. Additionally, while we did not measure the life length of our samples specifically, it might be that the commonly thought tendency of 3-aminoindoles toward oxidative breakdown is somewhat overrated.

3. Materials and Methods

3.1. General Information

NMR spectra, 1H, 13C, and 19F were measured in solutions of CDCl3 or DMSO-d6 on Bruker AVANCE-III HD instrument (at 400, 101, and 376 MHz, respectively). Residual solvent signals were used as internal standards in DMSO-d6 (2.50 ppm for 1H and 40.45 ppm for 13C nuclei) or CDCl3 (7.26 ppm for 1H and 77.16 ppm for 13C nuclei). HRMS spectra were measured on Bruker maXis impact (electrospray ionization in MeCN solutions, employing HCO2Na–HCO2H for calibration). IR spectra were measured on FT-IR spectrometer Shimadzu IRAffinity-1S equipped with an ATR sampling module. Spectral data are provided in the Supplementary Materials (S1–S42). Reaction progress, purity of isolated compounds, and Rf values were monitored with TLC on Silufol UV-254 plates. Column chromatography was performed on silica gel (32–63 μm, 60 Å pore size). Melting points were measured with Stuart SMP30 apparatus. All spirocyclic indoles 3 and indolinones 4, except 4ab and 4ad, were synthesized according to the previously reported procedures and were identical to those described [37]. All reagents and solvents were purchased from commercial vendors and used as received.

3.2. Preparation of 2-(5-Halide-3-Oxo-2-Phenylindolin-2-Yl)-2-Phenylacetonitrile 4ab, 4ad (General Procedure)

These compounds were prepared in analogy to the method described in [37]. A Weaton microreactor equipped with magnetic spin-vane and Mininert valve was charged with a mixture of (E)-(2-Nitrovinyl)benzene (150 mg, 1.0 mmol), corresponding 5-halide-2-phenyl-1H-indole (1.0 mmol), phosphorus acid (1.0 g), and formic acid (1.0 g). The mixture was vigorously stirred for 2 h at room temperature while it turned dark red and homogenized. Then, the mixture was poured into water (50 mL), and the formed crude spirane 3 precipitate was collected and washed with water (4 × 20 mL), dried, and dissolved in ethanol (4 mL). Triethylamine (102 mg, 1.0 mmol) was added, and the resulting solution was stirred at room temperature for 3 h. Crystalline precipitate of crude product was formed, which was collected and purified by preparative column chromatography on silica gel, eluting with ethyl acetate/hexane mixture (1:4).
2-(5-Fluoro-3-oxo-2-phenylindolin-2-yl)-2-phenylacetonitrile (4ab): yellowish solid, m.p. 200–202 °C, Rf 0.56 (EtOAc/hexane, 1:3, v/v). Yield 274 mg (0.80 mmol, 80%). 1H NMR (400 MHz, DMSO-d6) δ 8.13 (s, 1H), 7.55 (d, J = 7.7 Hz, 2H), 7.47 (t, J = 9.3 Hz, 1H), 7.36–7.19 (m, 9H), 7.16 (d, J = 9.5 Hz, 1H), 5.31 (s, 1H); 13C NMR (101 MHz, DMSO-d6) δ 198.7 (d, J = 3.7 Hz), 158.8, 155.6 (d, J = 237.0 Hz), 134.4, 131.5, 129.4 (2C), 128.5 (2C), 128.5, 128.4 (2C), 126.6, 126.3, 126.3 (2C), 118.9, 117.7 (d, J = 7.7 Hz), 114.1 (d, J = 7.7 Hz), 109.3 (d, J = 22.7 Hz), 74.3, 43.5; 19F NMR (376 MHz, DMSO-d6) δ −125.34; FTIR, vmax: 3351, 2361, 2248, 1698, 1363, 1621, 1558, 1487, 1462, 1330, 1260, 1202 cm−1; HRMS (ESI TOF) m/z calc’d. for C22H15FN2NaO [M+Na]+: 365.1061, found: 365.1056 (1.2 ppm).
2-(5-Chloro-3-oxo-2-phenylindolin-2-yl)-2-phenylacetonitrile (4ad): yellowish solid, m.p. 200–202 °C, Rf 0.49 (EtOAc/hexane, 1:2, v/v). Yield 215 mg (0.60 mmol, 60%). 1H NMR (400 MHz, DMSO-d6) δ 8.37 (s, 1H), 7.60–7.52 (m, 4H), 7.34–7.18 (m, 8H), 7.15 (d, J = 8.8 Hz, 1H), 5.33 (s, 1H); 13C NMR (101 MHz, DMSO-d6) δ 197.9, 160.3, 138.0, 134.3, 131.3, 129.4 (2C), 128.6 (2C), 128.52, 128.50, 128.4 (2C), 126.2 (2C), 123.7, 122.6, 118.8, 118.7, 114.3, 73.9, 43.5; FTIR, vmax: 3298, 2925, 2247, 1693, 1616, 1474, 1301, 1258, 1165, 1056 cm−1; HRMS (ESI TOF) m/z calc’d. for C22H15ClN2NaO+ [M+Na]+: 381.0765, found: 381.0775 (−2.5 ppm).

3.3. Preparation of 2-Aryl-1H-Indol-3-Amine 5aa–ai (General Procedure)

Method A: the corresponding 2-(3-Oxo-2-phenylindolin-2-yl)-2-phenylacetonitrile 4 (1.00 mmol) and 2 mL of hydrazine hydrate were charged in a G10 microwave vial. The vial was sealed and heated in an Anton Paar Monowave 300 microwave apparatus at 200 °C for 15 min. After completion of reaction, vial was opened, and the reaction mixture was concentrated in vacuo. Crude material purified by column chromatography (EtOAc/hexane, 1:4, v/v).
Method B: the corresponding 2,4′-Diaryl-4′H-spiro[indole-3,5′-isoxazole] 3 (1.00 mmol) and 2 mL of hydrazine hydrate were charged in a G10 microwave vial. The vial was sealed and heated in an Anton Paar Monowave 300 microwave apparatus at 200 °C for 15 min. After completion of reaction, vial was opened, and the reaction mixture was concentrated in vacuo. Crude material purified by column chromatography (EtOAc/hexane, 1:4, v/v).
2-Phenyl-1H-indol-3-amine (5aa): this compound [32] was prepared via both Method A (0.90 mmol, 90%) and Method B (0.80 mmol, 80%). White solid, m.p. 107–108 °C, lit. 110–111 °C, Rf 0.46 (EtOAc/hexane, 1:4, v/v). 1H NMR (400 MHz, DMSO-d6) δ 10.52 (s, 1H), 7.85–7.77 (m, 2H), 7.69 (d, J = 7.9 Hz, 1H), 7.44 (t, J = 7.8 Hz, 2H), 7.27 (d, J = 8.1 Hz, 1H), 7.20 (t, J = 7.4 Hz, 1H), 7.11–7.02 (m, 1H), 6.96–6.87 (m, 1H), 4.50 (s, 2H); 13C NMR (CPD) (101 MHz, DMSO-d6) δ 135.0, 133.6, 128.7 (2C), 125.1 (2C), 125.0, 123.0 (2C), 122.0, 118.9, 118.3, 117.4, 110.9; 13C NMR (DEPT135) (101 MHz, DMSO-d6) δ 128.5 (2C), 124.83 (2C), 124.78, 121.7, 118.1, 117.1, 110.6; 13C NMR (DEPTQ) (101 MHz, DMSO-d6) δ 135.0, 133.6, 128.7 (2C), 125.1 (2C), 125.0, 123.0 (2C), 122.0, 118.9, 118.3, 117.4, 110.9; FTIR, vmax: 3646, 3503, 3351, 3198, 1887, 1769, 1684, 1600, 1489, 1457, 1378, 1245 cm−1; HRMS (ESI TOF) m/z: calc’d for C14H13N2 [M+H]+: 209.1073, found 209.1069 (2.1 ppm).
5-Fluoro-2-phenyl-1H-indol-3-amine (5ab): this compound was prepared via Method A (0.77 mmol, 77%). White solid, m.p. 105–106 °C, Rf 0.33 (EtOAc/hexane, 1:3, v/v). 1H NMR (400 MHz, DMSO-d6) δ 10.58 (br. s, 1H), 7.86–7.72 (m, 2H), 7.51–7.40 (m, 3H), 7.25–7.17 (m, 2H), 6.87 (td, J = 9.2, 2.6 Hz, 1H), 4.47 (br. s, 2H); 13C{1H} NMR (101 MHz, DMSO-d6) δ 157.2, 154.9, 132.4 (d, J = 167.3 Hz), 128.7 (2C), 125.5, 125.2 (2C), 123.1 (d, J = 4.8 Hz), 122.8 (d, J = 9.9 Hz), 121.2, 111.7 (d, J = 9.5 Hz), 109.9 (d, J = 26.0 Hz), 103.0 (d, J = 23.8 Hz); 19F NMR (376 MHz, DMSO-d6) δ −126.28 (s); FTIR, vmax: 3629, 3371, 2993, 1922, 1771, 1688, 1522, 1248, 1174, 1061 cm−1; HRMS (ESI TOF) m/z: calc’d for C14H11FN2 [M+H]+: 227.0979, found 227.0976 (1.2 ppm).
5-Isopropyl-2-phenyl-1H-indol-3-amine (5ac): this compound was prepared via Method A (0.82 mmol, 82%). White solid, m.p. 77–79 °C, Rf 0.56 (EtOAc/hexane, 1:3, v/v). 1H NMR (400 MHz, DMSO-d6) δ 10.34 (br. s, 1H), 7.79 (dd, J = 8.5, 1.3 Hz, 2H), 7.53 (s, 1H), 7.45–7.39 (m, 2H), 7.20–7.13 (m, 2H), 6.96 (dd, J = 8.4, 1.8 Hz, 1H), 4.42 (s, 2H), 3.02–2.83 (m, 1H), 1.28 (s, 3H), 1.26 (s, 3H); 13C{1H} NMR (101 MHz, DMSO-d6) δ 137.5, 133.8, 133.7, 128.7 (2C), 125.0 (2C), 124.9, 123.0, 122.9, 121.3, 119.2, 115.00, 110.6, 33.7, 24.3 (2C); FTIR, vmax: 3381, 2950, 1922, 1827, 1655, 1524, 1459, 1240 cm−1; HRMS (ESI TOF) m/z: calc’d for C17H19N2 [M+H]+: 251.1543, found 251.1543 (1.2 ppm).
5-Chloro-2-phenyl-1H-indol-3-amine (5ad): this compound [2] was prepared via Method A (0.85 mmol, 85%). Yellow solid, m.p. 124–125 °C, lit. 125–127 °C, Rf 0.65 (EtOAc/hexane, 1:2, v/v). 1H NMR (400 MHz, DMSO-d6) δ 10.72 (s, 1H), 7.79 (d, J = 8.3 Hz, 3H), 7.45 (t, J = 7.6 Hz, 2H), 7.29–7.18 (m, 2H), 7.03 (dd, J = 8.6, 2.1 Hz, 1H), 4.58 (s, 2H); 13C NMR (101 MHz, DMSO-d6) δ 133.2, 133.1, 128.7 (2C), 125.5, 125.2 (2C), 123.8, 122.6, 121.9, 121.6, 120.5, 117.7, 112.3; FTIR, vmax: 3742, 3627, 3324, 3066, 2358, 1686, 1558, 1507, 1455, 1352 cm−1; HRMS (ESI TOF) m/z calc’d. for C14H12ClN2 [M+H]+: 243.0679, found: 243.0684 (1.8 ppm).
2-(Naphthalen-2-yl)-1H-indol-3-amine (5ae): this compound was prepared via Method A (0.70 mmol, 70%). Yellow solid, m.p. 183–186 °C, Rf 0.53 (EtOAc/hexane, 1:2, v/v). 1H NMR (400 MHz, DMSO-d6) δ 10.64 (s, 1H), 8.24 (d, J = 1.8 Hz, 1H), 8.04 (dd, J = 8.7, 1.8 Hz, 1H), 7.96 (d, J = 8.7 Hz, 1H), 7.93–7.85 (m, 2H), 7.71 (d, J = 7.9 Hz, 1H), 7.51 (ddd, J = 8.2, 6.8, 1.4 Hz, 1H), 7.45 (ddd, J = 8.0, 6.8, 1.3 Hz, 1H), 7.27 (d, J = 8.1 Hz, 1H), 7.08 (ddd, J = 8.1, 6.9, 1.2 Hz, 1H), 6.92 (ddd, J = 7.9, 6.9, 1.0 Hz, 1H), 4.69 (s, 2H); 13C NMR (101 MHz, DMSO-d6) δ 135.3, 133.6, 131.2, 130.9, 128.0, 127.7, 127.6, 126.4, 125.2, 124.2, 123.8, 122.9, 122.5, 122.2, 118.7, 118.4, 117.4, 110.9; FTIR, vmax: 3360, 3157, 3062, 2922, 2851, 1908, 1871, 1608, 1548, 1488 cm−1; HRMS (ESI TOF) m/z calc’d. for C18H15N2 [M+H]+: 259.1223, found: 259.1230 (2.8 ppm).
2-(4-Methoxyphenyl)-1H-indol-3-amine (5af): this compound [7] was prepared via Method A (0.68 mmol, 68%). White solid, m.p. 113–115 °C, lit. 108–110 °C, Rf 0.25 (EtOAc/hexane, 1:3, v/v). 1H NMR (400 MHz, DMSO-d6) δ 10.44 (br. s, 1H), 7.74 (d, J = 8.8 Hz, 2H), 7.62 (d, J = 7.8 Hz, 1H), 7.22 (d, J = 8.1 Hz, 1H), 7.06–6.99 (m, 3H), 6.89 (t, J = 7.4 Hz, 1H), 4.30 (br. s, 2H), 3.79 (s, 3H); 13C{1H} NMR (101 MHz, DMSO-d6) δ 157.1, 134.6, 126.6 (2C), 126.3, 123.3, 121.4, 121.3, 119.5, 117.9, 117.4, 114.2 (2C), 110.7, 55.2; FTIR, vmax: 3619, 3265, 2957, 1769, 1749, 1562, 1456, 1374, 1237, 1085 cm−1; HRMS (ESI TOF) m/z: calc’d for C15H15N2O [M+H]+: 239.1179, found 239.1184 (−2.3 ppm).
2-(p-Tolyl)-1H-indol-3-amine (5ag): this compound [7] was prepared via Method A (0.65 mmol, 65%). White solid, m.p. 149–151 °C, lit. 149–151 °C, Rf 0.61 (EtOAc/hexane, 1:2, v/v). 1H NMR (400 MHz, DMSO-d6) δ 10.45 (s, 1H), 7.68 (d, J = 7.9 Hz, 2H), 7.63 (d, J = 7.9 Hz, 1H), 7.23 (dd, J = 11.5, 7.8 Hz, 3H), 7.02 (ddd, J = 8.1, 6.9, 1.2 Hz, 1H), 6.88 (ddt, J = 8.1, 7.0, 1.1 Hz, 1H), 4.39 (s, 2H), 2.33 (s, 3H); 13C NMR (101 MHz, DMSO-d6) δ 134.8, 134.2, 130.8, 129.3 (2C), 125.0 (2C), 123.0, 122.3, 121.7, 119.1, 118.1, 117.3, 110.7, 20.8; FTIR, vmax: 3742, 3627, 3324, 3066, 2358, 1686, 1558, 1507, 1455, 1352 cm−1; HRMS (ESI TOF) m/z calc’d. for C15H15N2 [M+H]+: 223.1225, found: 223.1230 (2.1 ppm).
2-(5,6,7,8-Tetrahydronaphthalen-2-yl)-1H-indol-3-amine (5ah): this compound was prepared via Method A (0.9 mmol, 90%). Yellowish solid, m.p. 164–167 °C, Rf 0.63 (EtOAc/hexane, 1:2, v/v). 1H NMR (400 MHz, CDCl3) δ 7.64 (s, 1H), 7.54 (d, J = 7.8 Hz, 1H), 7.41 (dd, J = 7.9, 2.0 Hz, 1H), 7.36 (d, J = 1.8 Hz, 1H), 7.33 (d, J = 8.0 Hz, 1H), 7.21 (t, J = 7.0 Hz, 2H), 7.13 (t, J = 7.4 Hz, 1H), 3.52 (s, 2H), 2.86 (dt, J = 11.0, 5.1 Hz, 4H), 2.09–1.57 (m, J = 4.0 Hz, 4H); 13C NMR (101 MHz, CDCl3) δ 138.2, 135.9, 134.9, 130.2, 130.1, 126.7, 123.5, 123.4, 122.6, 121.4, 120.7, 119.1, 117.3, 111.0, 29.7, 29.3, 23.32, 23.29; FTIR, vmax: 3332, 2938, 2370, 1578, 1472, 1360, 1314, 1254, 1206, 1023 cm−1; HRMS (ESI TOF) m/z Calc’d. for C18H19N2 [M+H]+: 263.1544, found: 2639.1543 (−0.4 ppm).
2-(2,3-Dihydrobenzo[b][1,4]dioxin-6-yl)-1H-indol-3-amine (5ai): this compound was prepared via Method A (0.83 mmol, 83%). Pale yellow solid, m.p. 101–102 °C, Rf 0.15 (EtOAc/hexane, 1:3, v/v). 1H NMR (400 MHz, DMSO-d6) δ 10.41 (br. s, 1H), 7.61 (d, J = 7.9 Hz, 1H), 7.34–7.26 (m, 2H), 7.23–7.16 (m, 1H), 7.04–6.98 (m, 1H), 6.95–6.84 (m, 2H), 4.28 (br. s, 2H), 4.28–4.19 (m, 4H); 13C{1H} NMR (101 MHz, DMSO-d6) δ 143.6, 141.3, 134.6, 127.1, 123.2, 121.7, 121.5, 119.1, 118.5, 118.0, 117.4, 117.3, 113.8, 110.7, 64.3, 64.2; FTIR, vmax: 3619, 3265, 2957, 1769, 1749, 1562, 1456, 1374, 1237, 1085 cm−1; HRMS (ESI TOF) m/z: calc’d for C16H15N2O2 [M+H]+: 267.1128, found 267.1121 (2.8 ppm).
1-Methyl-2-phenyl-1H-indol-3-amine (19): this compound was prepared via Method A employing the corresponding N-methylated indolinone 18 (338 mg, 1.00 mmol) in a yield of 91 mg (0.41 mmol, 41%). Purification was performed by column chromatography (EtOAc/hexane, 1:4, v/v). The titled compound was obtained as white solid, Rf 0.36 (EtOAc/hexane, 1:4, v/v). 1H NMR (400 MHz, CDCl3) δ 7.61–7.49 (m, 3H), 7.49–7.42 (m, 2H), 7.41–7.37 (m, 1H), 7.33–7.29 (m, 1H), 7.27–7.24 (m, 1H), 7.14–7.09 (m, 1H), 3.61 (s, 3H), 2.89 (s, 2H); 13C{1H} NMR (101 MHz, CDCl3) δ 136.4, 131.5, 130.0, 129.0 (2C), 127.5, 125.2, 122.3, 121.8, 120.7, 118.7, 117.6, 109.5 (2C), 31.2; FTIR, vmax: 3459, 3356, 3321, 3066, 1651, 1531, 1511, 1431, 1311, 1167 cm−1; HRMS (ESI TOF) m/z: calc’d for C15H15N2Na [M+Na]+: 223.1230, found 223.1228 (1.0 ppm).
1-methyl-2-phenyl-1H-indole (J): the corresponding N-methylated indolinone 18 (0.30 mmol), hydroxylamine hydrochloride (0.3 mmol), triethylamine (0.3 mmol), and 2 mL xylene were charged in a G10 microwave vial. The vial was sealed and heated in an Anton Paar Monowave 300 microwave apparatus at 200 °C for 30 min. After completion of reaction vial was opened and the reaction mixture was concentrated in vacuo. Crude material purified by column chromatography (hexane) in a yield of 9 mg (0.045 mmol, 15%). The titled compound was obtained as light-yellow solid, m.p. 99–101 °C, Rf 0.71 (EtOAc/hexane, 1:10, v/v). 1H NMR (400 MHz, CDCl3) δ 7.71 (d, J = 7.9 Hz, 1H), 7.58 (d, J = 7.6 Hz, 2H), 7.53 (t, J = 7.5 Hz, 2H), 7.45 (dd, J = 16.2, 7.6 Hz, 2H), 7.32 (t, J = 7.6 Hz, 1H), 7.22 (t, J = 7.4 Hz, 1H), 6.64 (s, 1H), 3.80 (s, 3H); 13C NMR (101 MHz, CDCl3) δ 141.7, 138.4, 132.9, 129.5 (2C), 128.6 (2C), 128.1, 128.0, 121.8, 120.6, 120.0, 109.7, 101.7, 31.3.

3.4. Preparation of N-(2-Aryl-1H-Indol-3-yl)acetamide N-Ac 5aa and N-Ac 5ai (General Procedure)

The corresponding 2-(3-Oxo-2-arylindolin-2-yl)-2-phenylacetonitrile 4 (1.00 mmol), 1 mL of glacial acetic acid, and 2 mL of hydrazine hydrate were charged in a G10 microwave vial. The vial was sealed and heated in an Anton Paar Monowave 300 microwave apparatus at 200 °C for 15 min. After completion of reaction, vial was opened, and the reaction mixture was concentrated in vacuo. Crude material purified by column chromatography (EtOAc/hexane, 2:1, v/v).
N-(2-Phenyl-1H-indol-3-yl)acetamide (N-Ac 5aa): white solid, Rf 0.17 (EtOAc/hexane, 1:1, v/v). 1H NMR (400 MHz, DMSO-d6) δ 11.38 (s, 1H), 9.44 (s, 1H), 7.77 (d, J = 8.2 Hz, 2H), 7.48 (t, J = 7.7 Hz, 2H), 7.45–7.29 (m, 3H), 7.12 (t, J = 7.6 Hz, 1H), 7.00 (t, J = 7.5 Hz, 1H), 2.09 (s, 3H); 13C{1H} NMR (101 MHz, DMSO-d6) δ 169.7, 134.5, 131.7, 131.3, 128.8 (2C), 127.5, 126.7 (2C), 126.2, 122.0, 119.1, 118.5, 111.4, 110.8, 22.8; FTIR, vmax: 3435, 3265, 2918, 2863, 1648, 1458, 1377, 1338, 1242, 1191, 1148, 1113, 1075 cm−1; HRMS (ESI TOF) m/z: calc’d for C16H14N2ONa [M+Na]+: 273.0998, found 273.0995 (1.2 ppm).
N-(2-(2,3-Dihydrobenzo[b][1,4]dioxin-6-yl)-1H-indol-3-yl)acetamide (N-Ac 5ai): white solid, Rf 0.29 (EtOAc/hexane, 2:1, v/v). 1H NMR (400 MHz, DMSO-d6) δ 11.25 (s, 1H), 9.37 (s, 1H), 7.33 (d, J = 8.1 Hz, 1H), 7.30–7.23 (m, 3H), 7.09 (t, J = 6.9 Hz, 1H), 6.96 (d, J = 8.3 Hz, 2H), 4.28 (s, 4H), 2.08 (s, 3H); 13C{1H} NMR (101 MHz, DMSO-d6) δ 169.7, 143.5, 143.1, 134.3, 131.1, 126.3, 125.0, 121.7, 119.9, 119.0, 118.2, 117.4, 115.3, 111.2, 110.0, 64.3, 64.2, 22.9; FTIR, vmax: 3259, 1655, 1585, 1511, 1494, 1459, 1372, 1340, 1282, 1246, 1171, 1126, 1063 cm−1; HRMS (ESI TOF) m/z: calc’d for C18H16N2O3Na [M+Na]+: 331.1053, found 331.1046 (2.2 ppm).

4. Conclusions

A novel preparative method for the synthesis of diverse 3-aminoindoles 5 based on a microwave-assisted cascade reaction of 2-(3-oxoindolin-2-yl)-2-arylacetonitriles 4 with hydrazine hydrate was developed. Alternatively, the same transformation could also be carried out from 4′-phenyl-4′H-spiro[indole-3,5′-isoxazoles] 3. Considering that such spirocyclic indoles 3 could be obtained in a single step from commonly available indoles 1 and nitrostyrene 2, the overall sequence provides a very convenient and affordable route to generally not easily available 3-aminoindoles.

Supplementary Materials

The following supporting information that includes 1H, 13C, and 19F NMR spectral and HRMS charts can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28093657/s1.

Author Contributions

N.A.A. (Nicolai A. Aksenov)—conceptualization, supervision, data analysis, N.A.A. (Nikolai A. Arutiunov)—investigation, data analysis, I.A.K.—investigation, V.V.M.—investigation, D.A.A.—investigation, D.S.M.—investigation, A.M.Z.—investigation, E.A.C.—investigation, A.V.L.—writing (original draft, review, and editing), A.V.A.—funding acquisition, conceptualization, supervision, data analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the Russian Science Foundation (Grant number 21-73-20051, https://rscf.ru/en/project/21-73-20051/, accessed on 22 March 2023).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The supporting information includes NMR spectral and HRMS charts.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References

  1. Arzel, E.; Rocca, P.; Grellier, P.; Labaeïd, M.; Frappier, F.; Guéritte, F.; Gaspard, C.; Marsais, F.; Godard, A.; Quéguiner, G. New Synthesis of Benzo-δ-carbolines, Cryptolepines, and Their Salts: In Vitro Cytotoxic, Antiplasmodial, and Antitrypanosomal Activities of δ-Carbolines, Benzo-δ-carbolines, and Cryptolepines. J. Med. Chem. 2001, 44, 949–960. [Google Scholar] [CrossRef]
  2. Bahekar, R.H.; Jain, M.R.; Goel, A.; Patel, D.N.; Prajapati, V.M.; Gupta, A.A.; Jadav, P.A.; Patel, P.R. Design, synthesis, and biological evaluation of substituted-N-(thieno[2,3-b]pyridin-3-yl)-guanidines, N-(1H-pyrrolo[2,3-b]pyridin-3-yl)-guanidines, and N-(1H-indol-3-yl)-guanidines. Bioorg. Med. Chem. 2007, 15, 3248–3265. [Google Scholar] [CrossRef] [PubMed]
  3. Pews-Davtyan, A.; Tillack, A.; Schmöle, A.-C.; Ortinau, S.; Frech, M.J.; Rolfs, A.; Beller, M. A new facile synthesis of 3-amidoindole derivatives and their evaluation as potential GSK-3β inhibitors. Org. Biomol. Chem. 2010, 8, 1149–1153. [Google Scholar] [CrossRef]
  4. Nirogi, R.V.; Deshpande, A.D.; Kambhampati, R.; Badange, R.K.; Kota, L.; Daulatabad, A.V.; Shinde, A.K.; Ahmad, I.; Kandikere, V.; Jayarajan, P.; et al. Indole-3-piperazinyl derivatives: Novel chemical class of 5-HT6 receptor antagonists. Bioorg. Med. Chem. Lett. 2011, 21, 346–349. [Google Scholar] [CrossRef] [PubMed]
  5. Romagnoli, R.; Baraldi, P.G.; Sarkar, T.; Carrion, M.D.; Lopez-Cara, L.C.; Cruz-Lopez, O.; Preti, D.; Tabrizi, M.A.; Tolomeo, M.; Grimaudo, S.; et al. Synthesis and Biological Evaluation of 1-Methyl-2-(3′,4′,5′-trimethoxybenzoyl)-3-aminoindoles as a New Class of Antimitotic Agents and Tubulin Inhibitors. J. Med. Chem. 2008, 51, 1464–1468. [Google Scholar] [CrossRef] [PubMed]
  6. Dave, V.; Warnhoff, E.W. New reactions of 2-substituted indoles. Can. J. Chem. 1976, 54, 1020–1028. [Google Scholar] [CrossRef]
  7. Leijendekker, L.H.; Weweler, J.; Leuther, T.M.; Streuff, J. Catalytic Reductive Synthesis and Direct Derivatization of Unprotected Aminoindoles, Aminopyrroles, and Iminoindolines. Angew. Chem. Int. Ed. 2017, 56, 6103–6106. [Google Scholar] [CrossRef]
  8. Leijendekker, L.H.; Weweler, J.; Leuther, T.M.; Kratzert, D.; Streuff, J. Development, Scope, and Applications of Titanium(III)-Catalyzed Cyclizations to Aminated N-Heterocycles. Chem.-A Eur. J. 2019, 25, 3382–3390. [Google Scholar] [CrossRef]
  9. Zhang, G.; Lin, L.; Yang, K.; Wang, S.; Feng, Q.; Zhu, J.; Song, Q. 3-Aminoindole Synthesis from 2-Nitrochalcones and Ammonia or Primary Amines. Adv. Synth. Catal. 2019, 361, 3718–3722. [Google Scholar] [CrossRef]
  10. Nettekoven, M. Combinatorial synthesis of libraries of indole derivatives. Bioorg. Med. Chem. Lett. 2001, 11, 2169–2171. [Google Scholar] [CrossRef]
  11. Harjani, J.R.; Tang, A.X.; Norton, R.S.; Baell, J.B. A new methodology for the synthesis of 3-amino-1H-indole-2-carboxylates. Tetrahedron 2014, 70, 8047–8055. [Google Scholar] [CrossRef]
  12. Geng, X.; Wu, X.; Wang, C.; Zhao, P.; Zhou, Y.; Sun, X.; Wang, L.-J.; Guan, W.-J.; Wu, Y.-D.; Wu, A.-X. NaHS·nH2O-induced umpolung: The synthesis of 2-acyl-3-aminoindoles from aryl methyl ketones and 2-aminobenzonitriles. Chem. Commun. 2018, 54, 12730–12733. [Google Scholar] [CrossRef] [PubMed]
  13. Roy, S.; Roy, S.; Gribble, G.W. Efficient reductive acylation of 3-nitroindoles. Tetrahedron Lett. 2008, 49, 1531–1533. [Google Scholar] [CrossRef]
  14. Seong, C.M.; Park, C.M.; Choi, J.; Park, N.S. An efficient base-mediated intramolecular condensation of 2-(disubstituted amino)-benzonitriles to 3-aminoindoles. Tetrahedron Lett. 2009, 50, 1029–1031. [Google Scholar] [CrossRef]
  15. Gribble, G.; Saulnier, M.; Pelkey, E.; Kishbaugh, T.; Liu, Y.; Jiang, J.; Trujillo, H.; Keavy, D.; Davis, D.; Conway, S.; et al. Novel Indole Chemistry in the Synthesis of Heterocycles. Curr. Org. Chem. 2005, 9, 1493–1519. [Google Scholar] [CrossRef]
  16. Dang, Q.; Bai, X.; Xu, G.; Zheng, L. Total Synthesis of 4-Azaeudistomin Y1 and Analogues by Inverse-ElectronDemand Diels-Alder Reactions of 3-Aminoindoles with 1,3,5-Triazines. Synthesis 2013, 45, 743–752. [Google Scholar] [CrossRef]
  17. Gribble, G.W.; Roy, S. Convenient Synthesis of Masked Aminoindoles by Indium Mediated Ont-Pot Reductive Acylation of 3- and 2-Nitroindoles. Heterocycles 2006, 70, 51. [Google Scholar] [CrossRef]
  18. Prasad, P.K.; Kalshetti, R.G.; Reddi, R.N.; Kamble, S.P.; Sudalai, A. I2-mediated regioselective C-3 azidation of indoles. Org. Biomol. Chem. 2016, 14, 3027–3030. [Google Scholar] [CrossRef]
  19. Przheval’skii, N.M.; Skvortsova, N.S.; Magedov, I.V. New derivatives of 3-Aminoindole. Synthesis of 2-Aryl(Hetaryl)-3-(3,5-Dimethyl-1-Pyrazolyl)indoles. Chem. Heterocycl. Compd. 2004, 40, 1435–1441. [Google Scholar] [CrossRef]
  20. Przheval’Skii, N.M.; Skvortsova, N.S.; Magedov, I.V. Fischer Synthesis of 3-(N-Acylamino)-2-phenylindoles. Chem. Heterocycl. Compd. 2002, 38, 1055–1061. [Google Scholar] [CrossRef]
  21. Chernyak, D.; Chernyak, N.; Gevorgyan, V. Efficient and General Synthesis of 3-Aminoindolines and 3-Aminoindoles via Copper-Catalyzed Three-Component Coupling Reaction. Adv. Synth. Catal. 2010, 352, 961–966. [Google Scholar] [CrossRef] [PubMed]
  22. Pews-Davtyan, A.; Beller, M. Efficient and simple zinc-mediated synthesis of 3-amidoindoles. Org. Biomol. Chem. 2011, 9, 6331–6334. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, Z.; Zhang, Z.; Li, Z. Switchable Synthesis of 2-Methylene-3-aminoindolines and 2-Methyl-3-aminoindoles Using Calcium Carbide as a Solid Alkyne Source. Org. Lett. 2022, 24, 8067–8071. [Google Scholar] [CrossRef] [PubMed]
  24. Diao, P.-C.; Hu, M.-J.; Yang, H.-K.; You, W.-W.; Zhao, P.-L. Facile one-pot synthesis, antiproliferative evaluation and structure-activity relationships of 3-amino-1H-indoles and 3-amino-1H-7-azaindoles. Bioorg. Chem. 2019, 88, 102914. [Google Scholar] [CrossRef]
  25. Gribble, G.W. Synthesis and Reactions of Nitroindoles. In Progress in Heterocyclic Chemistry; Elsevier: Amsterdam, The Netherlands, 2020; Volume 31, pp. 83–117. [Google Scholar] [CrossRef]
  26. Abdel-Rahman, A.A.; El-Latif, M.M.A.; El-Essawy, F.A.; Barakat, Y.A. Synthesis and Antiviral Activity of 3-Aminoindole Nucleosides of 2-Acetamido-2-deoxy-D-glucose. Bull. Korean Chem. Soc. 2012, 33, 3417–3422. [Google Scholar] [CrossRef]
  27. Suvorov, N.N.; Velezheva, V.S.; Yarosh, A.V.; Erofeev, Y.V.; Kozik, T.N. Indole derivatives. Chem. Heterocycl. Compd. 1975, 11, 959–964. [Google Scholar] [CrossRef]
  28. Hooper, M.W.; Utsunomiya, M.; Hartwig, J.F. Scope and Mechanism of Palladium-Catalyzed Amination of Five-Membered Heterocyclic Halides. J. Org. Chem. 2003, 68, 2861–2873. [Google Scholar] [CrossRef]
  29. Muralidhar, P.; Kumar, B.S.; Nagaraju, K.; Maddila, S. A novel method for the synthesis of 3-aminoindoles using iodine and Cs2CO3 as catalyst. Chem. Data Collect. 2021, 33, 100731. [Google Scholar] [CrossRef]
  30. Ortiz, G.X., Jr.; Hemric, B.N.; Wang, Q. Direct and Selective 3-Amidation of Indoles Using Electrophilic N-[(Benzenesulfonyl)oxy]amides. Org. Lett. 2017, 19, 1314–1317. [Google Scholar] [CrossRef]
  31. Yamaguchi, T.; Yamaguchi, E.; Itoh, A. Cross-Dehydrogenative C–H Amination of Indoles under Aerobic Photo-oxidative Conditions. Org. Lett. 2017, 19, 1282–1285. [Google Scholar] [CrossRef]
  32. Chen, W.-L.; Li, K.; Liao, W.-C.; Liang, W.-F.; Qiu, P.-W.; Liang, C.; Su, G.-F.; Mo, D.-L. An iron(iii)-catalyzed dehydrogenative cross-coupling reaction of indoles with benzylamines to prepare 3-aminoindole derivatives. Green Chem. 2021, 23, 9610–9616. [Google Scholar] [CrossRef]
  33. Benkovics, T.; Guzei, I.A.; Yoon, T.P. Oxaziridine-Mediated Oxyamination of Indoles: An Approach to 3-Aminoindoles and Enantiomerically Enriched 3-Aminopyrroloindolines. Angew. Chem. Int. Ed. 2010, 49, 9153–9157. [Google Scholar] [CrossRef]
  34. Sokolovs, I.; Lubriks, D.; Suna, E. Copper-Catalyzed Intermolecular C–H Amination of (Hetero)arenes via Transient Unsymmetrical λ3-Iodanes. J. Am. Chem. Soc. 2014, 136, 6920–6928. [Google Scholar] [CrossRef] [PubMed]
  35. Aksenov, A.V.; Aksenov, N.A.; Aksenov, D.A.; Khamraev, V.F.; Rubin, M. Nitrostyrenes as 1,4-CCNO-dipoles: Diastereoselective formal [4+1] cycloaddition of indoles. Chem. Commun. 2018, 54, 13260–13263. [Google Scholar] [CrossRef]
  36. Aksenov, A.V.; Aksenov, D.A.; Arutiunov, N.A.; Aksenov, N.A.; Aleksandrova, E.V.; Zhao, Z.; Du, L.; Kornienko, A.; Rubin, M. Synthesis of Spiro[indole-3,5′-isoxazoles] with Anticancer Activity via a Formal [4 + 1]-Spirocyclization of Nitroalkenes to Indoles. J. Org. Chem. 2019, 84, 7123–7137. [Google Scholar] [CrossRef] [PubMed]
  37. Aksenov, A.V.; Aksenov, D.A.; Aksenov, N.A.; Aleksandrova, E.V.; Rubin, M. Preparation of Stereodefined 2-(3-Oxoindolin-2-yl)-2-Arylacetonitriles via One-Pot Reaction of Indoles with Nitroalkenes. J. Org. Chem. 2019, 84, 12420–12429. [Google Scholar] [CrossRef] [PubMed]
  38. Aksenov, A.V.; Aksenov, N.A.; Aleksandrova, E.V.; Aksenov, D.A.; Grishin, I.Y.; Sorokina, E.A.; Wenger, A.; Rubin, M. Direct Conversion of 3-(2-Nitroethyl)-1H-Indoles into 2-(1H-Indol-2-yl)Acetonitriles. Molecules 2021, 26, 6132. [Google Scholar] [CrossRef]
  39. Aksenov, A.V.; Aleksandrova, E.V.; Aksenov, D.A.; Aksenova, A.A.; Aksenov, N.A.; Nobi, M.A.; Rubin, M. Synthetic Studies toward 1,2,3,3a,4,8b-Hexahydropyrrolo[3,2-b]indole Core. Unusual Fragmentation with 1,2-Aryl Shift. J. Org. Chem. 2022, 87, 1434–1444. [Google Scholar] [CrossRef]
  40. Aksenov, A.V.; Kirilov, N.K.; Arutiunov, N.A.; Aksenov, D.A.; Kuzminov, I.K.; Aksenov, N.A.; Turner, D.N.; Rogelj, S.; Kornienko, A.; Rubin, M. Reductive Cleavage of 4′H-Spiro[indole-3,5′-isoxazoles] En Route to 2-(1H-Indol-3-yl)acetamides with Anticancer Activities. J. Org. Chem. 2022, 87, 13955–13964. [Google Scholar] [CrossRef] [PubMed]
  41. Aksenov, A.V.; Arutiunov, N.A.; Aksenov, D.A.; Samovolov, A.V.; Kurenkov, I.A.; Aksenov, N.A.; Aleksandrova, E.A.; Momotova, D.S.; Rubin, M. A Convenient Way to Quinoxaline Derivatives through the Reaction of 2-(3-Oxoindolin-2-yl)-2-phenylacetonitriles with Benzene-1,2-diamines. Int. J. Mol. Sci. 2022, 23, 11120. [Google Scholar] [CrossRef]
  42. Velezheva, V.S.; Brennan, P.J.; Marshakov, V.Y.; Gusev, D.V.; Lisichkina, I.N.; Peregudov, A.S.; Tchernousova, L.N.; Smirnova, T.G.; Andreevskaya, S.N.; Medvedev, A.E. Novel Pyridazino[4,3-b]indoles with Dual Inhibitory Activity against Mycobacterium tuberculosis and Monoamine Oxidase. J. Med. Chem. 2004, 47, 3455–3461. [Google Scholar] [CrossRef] [PubMed]
  43. Prasad, H.S.; Gowda, S.; Gowda, D.C. Facile Transfer Hydrogenation of Azo Compounds to Hydrazo Compounds and Anilines by Using Raney Nickel and Hydrazinium Monoformate. Synth. Commun. 2004, 34, 1–10. [Google Scholar] [CrossRef]
  44. Pasha, M.; Nanjundaswamy, H. Reductive fission of azoarenes to aminoarenes by aluminium/hydrazine hydrate. J. Chem. Res. 2004, 2004, 750–752. [Google Scholar] [CrossRef]
  45. Pasha, M.A.; NanjundaSwamy, H.M. Uncatalyzed Reductive Fission of Azoarenes to Aminoarene(s) by Hydrazine Hydrate. Synth. Commun. 2005, 35, 897–900. [Google Scholar] [CrossRef]
Scheme 1. Two-step C3-amination of 2-arylindoles with nitrostyrene and hydrazine hydrate.
Scheme 1. Two-step C3-amination of 2-arylindoles with nitrostyrene and hydrazine hydrate.
Molecules 28 03657 sch001
Scheme 2. Some transformations based on 2-(3-oxoindolin-2-yl)-2-arylacetonitriles 9.
Scheme 2. Some transformations based on 2-(3-oxoindolin-2-yl)-2-arylacetonitriles 9.
Molecules 28 03657 sch002
Scheme 3. The speculative approach to aminopyridazine 15.
Scheme 3. The speculative approach to aminopyridazine 15.
Molecules 28 03657 sch003
Scheme 4. The unexpected formation of 3-aminoindoles.
Scheme 4. The unexpected formation of 3-aminoindoles.
Molecules 28 03657 sch004
Scheme 5. Proposed mechanism of 3-aminoindoles formation with R3CH2NH2-type nucleophiles.
Scheme 5. Proposed mechanism of 3-aminoindoles formation with R3CH2NH2-type nucleophiles.
Molecules 28 03657 sch005
Scheme 6. Proposed mechanism of 3-aminoindoles formation with R3NHNH2-type nucleophiles.
Scheme 6. Proposed mechanism of 3-aminoindoles formation with R3NHNH2-type nucleophiles.
Molecules 28 03657 sch006
Scheme 7. Plausible mechanism of formation of 1-methyl-2-phenylindole J.
Scheme 7. Plausible mechanism of formation of 1-methyl-2-phenylindole J.
Molecules 28 03657 sch007
Scheme 8. Unprotected 2-aryl-3-aminoindoles prepared by described herein procedure.
Scheme 8. Unprotected 2-aryl-3-aminoindoles prepared by described herein procedure.
Molecules 28 03657 sch008
Scheme 9. Some examples of N-acylated 3-aminoindoles 5.
Scheme 9. Some examples of N-acylated 3-aminoindoles 5.
Molecules 28 03657 sch009
Table 1. Reaction of 4aa with amino binucleophiles RXHNH2 under different conditions.
Table 1. Reaction of 4aa with amino binucleophiles RXHNH2 under different conditions.
Molecules 28 03657 i001
#Binucleophile, RXHNH2ConditionsYield 5aa(Isolated), %
1Ethylenediamine1-butanol, MW 200 °C, 0.5 h37
2Propane-1,2-diamine1-butanol, MW 200 °C, 0.5 h41
3Formamidine acetate1-butanol, MW 185 °C, 1 hdecomposed
4ThiosemicarbazideXylene, MW 180 °C, 0.5 h15
52-hydrazinylpyridineXylene, MW 160 °C, 0.5 h20
6PhenylhydrazineXylene, MW 200 °C, 0.5 h35
7TosylhydrazineXylene, MW 200 °C, 0.5 h39
8BenzohydrazideXylene, MW 200 °C, 0.5 h26
9Hydrazine hydrateneat, MW 200 °C, 15 min90
10HydroxylamineXylene, MW 200 °C, 0.5 hn.d.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Aksenov, N.A.; Arutiunov, N.A.; Kurenkov, I.A.; Malyuga, V.V.; Aksenov, D.A.; Momotova, D.S.; Zatsepilina, A.M.; Chukanova, E.A.; Leontiev, A.V.; Aksenov, A.V. A Two-Step Synthesis of Unprotected 3-Aminoindoles via Post Functionalization with Nitrostyrene. Molecules 2023, 28, 3657. https://doi.org/10.3390/molecules28093657

AMA Style

Aksenov NA, Arutiunov NA, Kurenkov IA, Malyuga VV, Aksenov DA, Momotova DS, Zatsepilina AM, Chukanova EA, Leontiev AV, Aksenov AV. A Two-Step Synthesis of Unprotected 3-Aminoindoles via Post Functionalization with Nitrostyrene. Molecules. 2023; 28(9):3657. https://doi.org/10.3390/molecules28093657

Chicago/Turabian Style

Aksenov, Nicolai A., Nikolai A. Arutiunov, Igor A. Kurenkov, Vladimir V. Malyuga, Dmitrii A. Aksenov, Daria S. Momotova, Anna M. Zatsepilina, Elizaveta A. Chukanova, Alexander V. Leontiev, and Alexander V. Aksenov. 2023. "A Two-Step Synthesis of Unprotected 3-Aminoindoles via Post Functionalization with Nitrostyrene" Molecules 28, no. 9: 3657. https://doi.org/10.3390/molecules28093657

Article Metrics

Back to TopTop