Next Article in Journal
Bioactive Dimeric Abietanoid Peroxides from the Bark of Cryptomeria japonica
Next Article in Special Issue
Development of One Pot Strategy for Hyper Production and In Vivo Evaluation of Lovastatin
Previous Article in Journal
Synthesis and Biological Evaluation of NH2-Sulfonyl Oseltamivir Analogues as Influenza Neuraminidase Inhibitors
Previous Article in Special Issue
The Role Ionic Liquid [BMIM][PF6] in One-Pot Synthesis of Tetrahydropyran Rings through Tandem Barbier–Prins Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Metal-Free Air Oxidation in a Convenient Cascade Approach for the Access to Isoquinoline-1,3,4(2H)-triones

Dipartimento di Chimica e Biologia “A. Zambelli”, Università degli studi di Salerno, Via Giovanni Paolo II, 84084-Fisciano, Salerno, Italy
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(11), 2177; https://doi.org/10.3390/molecules24112177
Submission received: 14 May 2019 / Revised: 6 June 2019 / Accepted: 7 June 2019 / Published: 10 June 2019
(This article belongs to the Special Issue Recent Advances in Cascade Reactions and Related One-Pot Processes)

Abstract

:
Herein we describe a very useful application of the readily available trifunctional aromatic ketone methyl-2-(2-bromoacetyl)benzoate in reactions with primary amines. An unexpected in situ air oxidation that follows a cascade process allowed the access to a series of isoquinoline-1,3,4(2H)-triones, a class of heterocyclic compounds of great interest containing an oxygen-rich heterocyclic scaffold. A modification of the original protocol, utilizing a Staudinger reaction in the presence of trimethylphosphine, was necessary for the synthesis of Caspase inhibitor trione with free NH group.

Graphical Abstract

1. Introduction

The interest of academia and industry toward one-pot reactions has notably increased during the last two decades because of the environmentally benign conditions deriving from the high level of atom and step economy [1]. These transformations, which also include cascade, tandem, and domino reactions, generally refer to a process involving two or more consecutive reactions, without the isolation of the intermediates. The development of effective one-pot reactions is related to the rational design of multifunctional starting materials [2]. Within this context, we have recently investigated the reactivity as electrophiles in cascade reactions of easily accessible bifunctional aromatic ketones like 2-acylbenzonitriles [3,4,5], developing a particularly convenient process for the synthesis of highly functionalized isoindolinones [3,4,5], a valuable class of heterocyclic compounds [6]. Continuing the efforts to explore the chemistry of multifunctional ketones, we were interested to investigate the reactivity of readily available methyl 2-(2-bromoacetyl)benzoate 1 with primary amines (Scheme 1). There are relatively few reports on the applications of methyl 2-(2-bromoacetyl)benzoate, which consist of displacement reactions with thiols for the synthesis of anticancer agents 1,4-disubstituted phthalazines [7,8] or the preparation of Gd(III)-complexes involving reactions with secondary amines [9]. In this project, we have focused on the possibility to obtain in a single pot procedure 2,3-dihydroisoquinoline-1,4-diones 2 and to study the second reactivity for instance in oxidation reactions, which can allow the access to 2,3-dihydroisoquinoline-1,3,4-triones 3 (Scheme 1).
2,3-Dihydroisoquinoline-1,3,4-triones 3 are highly appealing for the number of biological activities as caspase inhibitors [10,11,12], nerve protectors to treat various neurodegenerative diseases, especially Alzheimer’s disease apoplexy and ischemic brain injuries [12,13]. Amino-substituted derivatives have been found to be active as herbicides due to their efficient redox mediation of photosystem I and their activity was found to be greater than that of the parent trione (Figure 1) [14,15,16,17]. The semicarbazide derivative prepared from phthalonimide shows good binding affinity toward oxytocin receptors (Ki = 1.6 nM) [18]. In addition, these compounds are particularly useful as intermediates in the synthesis of biologically active alkaloids such as benzo[c]phenanthridine alkaloids or nuevamine [19,20]. In contrast to their fascinating biological profiles, the methods for synthesis of these compounds are relatively limited, usually requiring multi-step sequences employing harsh conditions or harmful reagents with tedious purifications of the intermediates or show a limited substrate scope (Scheme 1).
One of the first synthesis approach was carried out by the oxidation of 3,4-dihydroisoquinolin-1,3-dione with RuO4 as a catalyst and NaIO4 as the stoichiometric oxidant (Scheme 1a) [21]. Later modifications included the oxidation by SeO2 (Scheme 1a) [10], tetraphenylporphyrin sensitized photooxygenation (Scheme 1a) [22,23] or the oxidation of C5-substituted isoquinolin-1-one by K2Cr2O7 (Scheme 1b) [10]. Phthalimides can be transformed into isoquinolinetriones after alkylation with chloroacetone and ring expansion in the presence of excess of sodium methoxide to give 4-hydroxyisocarbostyrils. These intermediates in a suitable solvent in the presence of oxygen undergo to an oxidative deacylation (Scheme 1c) [12]. Some isoquinolinetriones can be prepared from the corresponding secondary benzamides, as shown in Scheme 1d: The reactions of appropriate benzamides and oxalyl chloride lead to N-aroyloxamoyl chlorides. These isolable intermediates undergo cyclization upon heating when the group X is an activating group such as an alkoxy substituent [24]. Besides the above reactions involving the oxidation of de-aromatized isoquinolines, isoquinoline-1,3,4-trione can also be prepared by the Beckmann rearrangement or the azido-Schmidt reaction of ninhydrin, using hydroxylamine hydrochloride [25] or trimethylsilyl azide [26], respectively. Recently, a more direct access to dihydroisoquinoline-1,3,4(2H)-triones has been reported using I2/TBHP as oxidant, but it suffers from a limited substrate scope and harsh reaction conditions (Scheme 1e) [27]. Even though isoquinoline-1,3-diones are known compounds both as intermediates in the synthesis of 2,3-dihydroisoquinoline-1,3,4-triones and bioactive molecules [28,29,30,31,32,33], to the best of our knowlwdge, the analogues 2,3-dihydroisoquinoline-1,4-diones 2 have never been reported in literature. Therefore, the development of an efficient method to prepare 2,3-dihydroisoquinolines-1,4-diones from readily available sources and investigation of their reactivity would be highly desirable.

2. Results

In order to develop a direct access to 2,3-dihydroisoquinolines-1,4-diones 2, we treated 2-(2-bromoacetyl)benzoate 1 with one equivalent of benzyl amine, used as model nucleophile, at room temperature in dry CH3CN (Table 1). However, under the conditions reported in Entries 1 and 2 of Table 1, the unreacted starting materials were recovered. The conversion was significantly improved in the presence of one equivalent of diisopropylethylamine (DIPEA) at 50 °C (Entry 3–5). In this case, instead of the expected 2,3-dihydroisoquinoline-1,4-dione 2, we isolated the respective 2,3-dihydroisoquinoline-1,3,4-trione 3a in a very good 71% yield (Entry 5). The structure of the obtained product was clearly attributed by single-crystal X-ray diffraction, MS spectrum, and comparison of 1H-NMR spectrum with literature [27]. The X-ray molecular structure is reported in Figure 2 (for details check Supplementary Materials).
In a control experiment performed under nitrogen atmosphere at 50 °C (Entry 6), we observed the formation of a mixture of products, and only a tiny amount of 3a was isolated after chromatography. Since 1H-NMR performed on the crude did not highlight the presence of 3a, its formation was probably due to slow air oxidation during purification and isolation procedures (Entry 6). We also tested other bases. However, either Et3N or inorganic bases like NaHCO3 and K2CO3 were less effective, leading to decomposition products in a larger extent (Entries 7–11).
Then, under the optimized reaction conditions, we briefly analyzed the scope of the method using a series of aliphatic and benzyl primary amines (Table 2). In all the cases, also with the highly volatile methylamine or ethylamine, added as THF solutions, as well as in the presence of benzyl amines substituted with both electron-withdrawing or -donating groups, we obtained the 2,3-dihydroisoquinoline-1,3,4-triones in good yields. Only in the presence of the less nucleophilic aniline, even though after 18 h at 50 °C the starting materials disappeared, we recovered a complex mixture of products (Entry 10).
We also investigated the reaction of 1 with ammonia, used as 0.5 M solution in dioxane in order to developed one-pot protocol in the synthesis of the 3-caspase inhibitor [12]. The reaction, as summarized in Scheme 2A, was sluggish. The expected product was obtained only in 30% yield together with a series of not well identified intermediates and decomposition products. To this purpose, we modified the original protocol, reacting 1a with NaN3 and the crude derivative was subjected to Staudinger reaction [34] in the presence of PMe3 [35]. Even though the isolation and purification of the azide intermediate was necessary in this case, the reduction, cyclization, and air oxidation were carried out in a single pot reaction, simply changing N2 with air at the end of the cascade reaction and reacting for further 18 h at 50 °C. This furnished the expected 3j in a satisfying 62% yield.
On the basis of the obtained results and the control experiments, we can confidently propose the mechanism as described in Scheme 3. In particular, we believe that the first steps of the process involve a nucleophilic displacement of bromine and then lactamization. In this sequence of reactions, one equivalent of hindered DIPEA is necessary to neutralize the formed HBr, while the less-hindered Et3N could give competing nucleophilic displacement. Once formed the supposed 2,3-dihydroisoquinoline-1,4-diones intermediate 2, this is subject to in situ air auto-oxidation probably via enol formation 5 (Scheme 4).
Aerobic oxidations without the use of any reagent or metal catalyst are quite rare, but two notable examples have been recently reported for the aerobic α-oxidation of nitrogen containing heterocycles [36,37]. In particular, Foss et al. described an isoindolinone synthesis by selective metal-free aerobic oxidation of isoindolines [36]. In addition, Tang et al. reported an aerobic hydroxylation of isoquinoline-1,3-diones to 4-hydroxy-isoquinoline-1,3-diones by air oxidation via enol formation [37]. Supported by these studies, the enol 5 may react with oxygen in the air, with a following protonation to give hydroperoxide 6. Then, water elimination furnishes the 2,3-dihydroisoquinoline-1,3,4-triones 3.

3. Materials and Methods

3.1. General Information

Reactions were performed using commercially available compounds without further purification and analytical grade solvents. All the reactions were monitored by thin layer chromatography (TLC) on precoated silica gel plates (0.25 mm) and visualized by fluorescence quenching at 254 nm. Flash chromatography was carried out using silica gel 60 (70–230 mesh, Merck, Darmastdt, Germany). The NMR spectra were recorded on Bruker (Rheinstetten, Germany) 400 and 300 spectrometers (400 MHz, 1H, 100 MHz, 13C, 300 MHz, 1H, 75 MHz, 13C). Spectra were referenced to residual CHCl3 (7.26 ppm, 1H, 77.00 ppm, 13C). The following abbreviations are used to indicate the multiplicity in NMR spectra: s—singlet, d—doublet, t—triplet, q—quartet, dd—double doublet, ddd—doublet of doublet of doublet, m—multiplet, bs—broad signal. Coupling constants (J) are quoted in hertz. Yields are given for isolated products showing one spot on a TLC plate and no impurities detectable in the NMR spectrum. FTIR spectra were recorded as thin films on KBr plates using Bruker (Rheinstetten, Germany) VERTEX 70 spectrometer and absorption maxima are reported in wavenumber (cm−1). High-resolution mass spectra (HRMS) were acquired using a Bruker solariX XR Fourier transform ion cyclotron resonance mass spectrometer (Bruker Daltonik GmbH, Bremen, Germany) equipped with a 7T refrigerated actively shielded superconducting magnet. The samples were ionized in a positive ion mode using an electrospray (ESI) ionization source. Methyl 2-(bromoacetyl)benzoate was prepared according to the literature procedure [7].

3.2. General Procedure for Isoquinoline-1,3,4(2H)-triones Synthesis

Methyl 2-(bromoacetyl)benzoate (0.24 mmol, 1 equiv.) was dissolved in dry CH3CN (2 mL) and DIPEA (1 equiv.) was added followed by the addition of amine (1 equiv.). The mixture was vigorously stirred under air at 50 °C overnight in a round bottomed flask with a rubber stopper and a needle. The solvent was removed under vacuum and the crude was purified by flash column chromatography (Hexane 90/Ethyl acetate 10) using silica gel deactivated with 1% Et3N.

3.3. Product Characterization

2-benzylisoquinoline-1,3,4(2H)-trione (3a) [10,21,27,38]: Light yellow solid, 71% yield (45 mg). Mp. 190– 191 °C. Data in accordance with literature. 1H-NMR (400 MHz, CDCl3) δ 8.36 (dd, J = 1.0, 7.8 Hz, 1H), 8.20 (dd, J = 7.6, 1.0 Hz, 1H), 7.88 (td, J = 1.4, 7.6 Hz, 1H), 7.81 (dt, J = 1.3, 7.6 Hz, 1H), 7.51–7.49 (m, 2H), 7.35–7.27 (m, 3H), 5.24 (s, 2H). 13C-NMR (100 MHz, CDCl3) δ 174.6, 162.0, 156.8, 136.1, 135.9, 134.4, 130.8, 129.8, 129.7, 129.3, 128.6, 127.7, 127.3, 44.3. HRMS (ESI): calcd. for [M + H]+ C16H12NO3: 266.0812, found: 266.0814.
2-(4-chlorobenzyl)isoquinoline-1,3,4(2H)-trione (3b) [27]: Light yellow solid, 73% yield (52 mg). Mp. 187–188 °C. Data in accordance with literature. 1H-NMR (300 MHz, CDCl3) δ 8.35 (d, J = 7.7 Hz, 1H), 8.22 (d, J = 7.6 Hz, 1H), 7.91 (td, J = 1.2, 7.6 Hz, 1H), 7.84 (td, J = 1.2, 7.6 Hz, 1H), 7.46 (d, J = 8.4 Hz, 2H), 7.28 (d, J = 8.4 Hz, 2H), 5.20 (s, 2H). 13C-NMR (75 MHz, CDCl3) δ 174.7, 162.4, 157.3, 136.7, 134.5, 134.0, 134.3, 131.2, 130.9, 130.8, 129.0, 128.8, 128.1, 43.6. HR-MS (ESI): calcd. for [M + H]+ C16H11 ClNO3: 300.0422, found: 300.0427.
2-(4-methoxybenzyl)isoquinoline-1,3,4(2H)-trione (3c) [27]: Light yellow solid, 75% yield (57 mg). Mp. 163–164 °C. Data in accordance with literature. 1H-NMR (CDCl3, 300 MHz) δ 8.33 (dd, J= 7.4 Hz, 1H), 8.17 (dd, J = 7.5 Hz, 1H), 7.91–7.82 (m, 2H), 7.47 (d, J = 8.5 Hz, 2H), 6.82 (d, J = 8.3 Hz, 2H), 5.15 (s, 2H), 3.76 (s, 3H). 13C-NMR (CDCl3, 100 MHz) δ 174.8, 162.3, 159.5, 156.8, 135.9, 134.3, 131.3, 130.6, 130.1, 129.7, 128.1, 127.6, 113.8, 55.1, 43.6. HR-MS (ESI): calcd. for [M + Na]+ C17H13NO4Na: 318.0736, found: 318.0733.
2-(2-fluorobenzyl)isoquinoline-1,3,4(2H)-trione (3d) [12]: Light yellow solid, 69% yield (47 mg). Mp. 186–187 °C. Data in accordance with literature. 1H-NMR (CDCl3, 400 MHz) δ 8.36 (dd, J = 0.8, 7.8 Hz, 1H), 8.25 (dd, J = 0.8, 7.8 Hz, 1H), 7.94–7.82 (m, 2H), 7.34–7.31 (m, 1H), 7.29–7.25 (m, 1H), 7.09–7.04 (m, 2H), 5.36 (s, 2H). 13C-NMR (CDCl3, 75 MHz) δ 174.7, 162.0 (J = 240 Hz), 161.9, 156.9, 136.3, 134.5, 130.8, 130.1 (J = 3.0 Hz), 130.0, 129.7, 129.6 (J = 9.0 Hz), 127.9, 124.1 (J = 3.0 Hz), 122.6 (J = 9.0 Hz), 115.6 (J = 9.0 Hz), 38.3 (J = 3.0 Hz). HR-MS (ESI): calcd. for [M + H]+ C16H11FNO3: 284.1707, found: 284.1710.
2-(3-nitrobenzyl)isoquinoline-1,3,4(2H)-trione (3e) [27]: Light yellow solid, 69% yield (51 mg). Mp. 186–187 °C. Data in accordance with literature. 1H-NMR (CDCl3, 400 MHz) δ 8.38 (t, J = 7.7 Hz, 2H), 8.24 (d, J = 7.6 Hz, 1H), 8.16 (J = 8.0 Hz, 1H), 7.97–7.94 (m, 1H), 7.89–7.82 (m, 2H), 7.52 (t, J = 8.0 Hz, 1H), 5.34 (s, 2H). 13C-NMR (100 MHz, CDCl3) δ 174.7, 162.3, 157.1, 148.8, 137.5, 136.4, 135.2, 135.9, 131.3, 130.3, 129.9, 129.8, 128.2, 124.5, 123.4, 43.8. HR-MS (ESI): calcd. for [M + H]+ C16H11N2O5: 311.0662, found: 311.0659.
2-Methylisoquinoline-1,3,4(2H)-trione (3f) [10,12,21,38]: Light yellow solid, 72% yield (32 mg). Mp. 186–187 °C. Data in accordance with literature. 1H-NMR (CDCl3, 300 MHz) δ 8.36 (dd, J = 1.3, 7.6 Hz, 1H), 8.23 (dd, J = 1.3, 7.6 Hz, 1H), 7.94–7.81 (m, 2H), 3.50 (s, 3H). 13C-NMR (CDCl3, 100 MHz) δ 174.5, 162.4, 157.3, 136.20, 134.4, 130.6, 129.8, 129.7, 127.8, 27.5. HR-MS (ESI): calcd. for [M + H]+ C10H8NO3: 190.0499, found: 190.0492.
2-Ethylisoquinoline-1,3,4(2H)-trione (3g) [14,21,38,39]: Light yellow solid, 70% yield (34 mg). Mp. 101–102 °C. Data in accordance with literature. 1H-NMR (CDCl3, 300 MHz) δ 8.38 (dd, J = 0.6, 5.8 Hz, 1H), 8.24 (dd, J = 0.8, 5.8 Hz, 1H), 7.94–7.83 (m, 2H), 4.15 (q, J = 7.1 Hz, 2H), 1.28 (t, J = 7.1 Hz, 3H). 13C-NMR (CDCl3, 100 MHz) δ 174.6, 161.8, 156.7, 135.9, 134.3, 130.7, 129.9, 129.7, 127.7, 37.3, 13.1. HR-MS (ESI): calcd. for [M + H]+ C11H10NO3: 204.0655, found: 204.0658.
2-Butylisoquinoline-1,3,4(2H)-trione (3h) [14,39]: Light yellow solid, 73% yield (40 mg). Mp. 60–61 °C. Data in accordance with literature. 1H-NMR (CDCl3, 300 MHz) δ 8.33 (d, J = 7.9 Hz, 1H), 8.20 (d, J = 7.9 Hz, 1H), 7.98–7.82 (m, 2H), 4.11 (t, J = 7.4 Hz, 2H), 1.74–1.68 (m, 2H), 1.66–1.41 (m, 2H), 1.03 (t, J = 7.4 Hz, 3H). 13C-NMR (CDCl3, 100 MHz) δ 174.7, 162.1, 156.9, 135.9, 134.3, 130.8, 129.9, 129.7, 127.7, 41.0, 29.9, 20.2, 13.7. HR-MS (ESI): calcd. for [M + H]+ C13H14NO3: 232.0968, found: 232.0963.
2-Allylisoquinoline-1,3,4(2H)-trione (3i) [10,12]: Light yellow solid, 72% yield (37 mg). Mp. 180–181 °C. Data in accordance with literature. 1H-NMR (CDCl3, 400 MHz) δ 8.35 (dd, J = 0.7, 7.8 Hz, 1H), 8.22 (dd, J = 0.8, 7.7 Hz, 1H), 7.94–7.83 (m, 2H), 5.96–5.86 (m, 1H), 5.38 (d, J = 17 Hz, 1H), 5.26 (d, J = 10 Hz, 1H), 4.66 (d, J = 6.1 Hz, 2H). 13C-NMR (CDCl3, 75 MHz) δ 174.7, 162.1, 156.9, 136.3, 134.7, 134.5, 131.0, 130.1, 130.0, 128.1, 119.4, 43.4. HR-MS (ESI): calcd. for [M + H]+ C10H12NO3: 216.0655, found: 216.0657.

3.4. General Procedure for Isoquinoline-1,3,4(2H)-trione Synthesis via Staudinger Reaction

Methyl 2-(bromoacetyl)benzoate (100 mg, 0.39 mmol, 1 equiv.) and NaN3 (1.5 equiv.) were vigorously stirred at room temperature overnight. The suspension was diluted with MTBE and solid was filtered off. The solvent was removed under vacuum and the crude was purified by flash column chromatography on silica gel (Hexane 95/Ethyl acetate 5). To a solution of azide 4 (84 mg, 0.38 mmol, 1 equiv.) in THF/H2O (2 mL/200 μL) under nitrogen atmosphere, trimethylphosphine (1M in THF, 456 μL, 1.2 equiv) was added and the mixture was stirred for 4 h at room temperature. Then, the mixture was stirred under air at 50 °C overnight utilizing a rubber stopper with a needle. After evaporation of the solvent, the crude was taken up with dichloromethane and washed with water. Purification by chromatography (Ethyl acetate 3/CHCl3 7) using silica gel deactivated wih 1% Et3N gave 3j as yellow solid.
Methyl 2-(2-azidoacetyl)benzoate (4): Pale oil, 98% yield (84 mg). 1H-NMR (CDCl3, 400 MHz) δ 7.99 (d, J = 8.0 Hz, 1H), 7.61 (t, J = 8.0 Hz, 1H), 7.56 (J = 8.0 Hz, 1H), 7.30 (d, J = 8.0 Hz, 1H), 4.25 (s, 2H), 3.90 (s, 3H). 13C-NMR (CDCl3, 100 MHz) δ 200.2, 166.3, 140.7, 132.9, 130.4, 130.2, 127.9, 126.3, 57.8, 52.8. HR-MS (ESI): calcd. for [M + H]+ C10H10N3O3: 219.0644, found: 219.0638.
Isoquinoline-1,3,4(2H)-trione (3j) [12,21,38]: Yellow solid, 62% yield (41 mg). Mp. 220–221 °C. Data in accordance with literature. 1H-NMR (DMSO-d6, 400 MHz) δ 11.98 (br, 1H), 8.14 (d, J = 7.9 Hz, 1H), 8.07 (d, J = 7.6 Hz, 1H), 7.95–7.88 (m, 2H). 13C-NMR (DMSO-d6, 75 MHz) δ 174.8, 162.7, 157.4, 136.5, 134.8, 131.2, 130.1, 129.7, 128.1. HR-MS (ESI): calcd. for [M + H]+ C9H6NO3: 176.0342, found: 176.0346.

4. Conclusions

In conclusion, we report simple and synthetically useful protocols for the preparation of 2,3-dihydroisoquinoline-1,3,4-triones by one-pot cascade reactions of methyl 2-bromoacetyl)benzoate with primary amines followed by in situ oxidation promoted by air, without the use of any other reagent or catalyst. This transformation proved to be of general applicability with respect to aliphatic or benzylic amines, while aromatic amines like aniline lead to decomposition products. A suitable modification of the original protocol utilizing a Staudinger reaction, allowed the synthesis of the NH-unsubstituted caspase inhibitor trione. On the basis of the present work, other studies are in course in order to analyze the applicability of the concepts herein developed and in particular of the metal-free air oxidation on related multifunctional substrates.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/24/11/2177/s1, Figure S1: The X-ray molecular structure of compound 3a and X-Ray data and NMR spectra of all the compounds.

Author Contributions

Methodology and investigation, A.D.M.; X-ray structural analysis, C.T.; writing—original draft preparation and supervision, A.M.; writing—review and editing, A.M.

Funding

This research received no external funding.

Acknowledgments

This work was supported by MIUR and University of Salerno.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chanda, T.; Zhao, J.C.-G. Recent Progress in Organocatalytic Asymmetric Domino Transformations. Adv. Synth. Catal. 2018, 360, 2–79. [Google Scholar] [CrossRef]
  2. Ravichandiran, P.; Lai, B.; Gu, Y. Aldo-X Bifunctional Building Blocks for the Synthesis of Heterocycles. Chem. Rec. 2017, 17, 142–183. [Google Scholar] [CrossRef] [PubMed]
  3. Di Mola, A.; Di Martino, M.; Capaccio, V.; Pierri, G.; Palombi, L.; Tedesco, C.; Massa, A. Synthesis of 2-Acetylbenzonitriles and Their Reactivity in Tandem Reactions with Carbon and Hetero Nucleophiles: Easy Access to 3,3-Disubstituted Isoindolinones. Eur. J. Org. Chem. 2018, 2018, 1699–1708. [Google Scholar] [CrossRef]
  4. Romano, F.; Di Mola, A.; Palombi, L.; Tiffner, M.; Waser, M.; Massa, A. Synthesis and Organocatalytic Asymmetric Nitro-aldol Initiated Cascade Reactions of 2-Acylbenzonitriles Leading to 3,3-Disubstituted Isoindolinones. Catalysts 2019, 9, 327. [Google Scholar] [CrossRef]
  5. Di Mola, A.; Macchia, A.; Tedesco, C.; Pierri, G.; Palombi, L.; Filosa, R.; Massa, A. Synthetic Strategies and Cascade Reactions of 2-Cyanobenzophenones for the Access to Diverse 3,3-Disubstituted Isoindolinones and 3-Aryl-3-Hydroxyisoindolinones. ChemistrySelect 2019, 4, 4820–4826. [Google Scholar] [CrossRef]
  6. Di Mola, A.; Palombi, L.; Massa, A. An overview on asymmetric synthesis of 3-substituted indolinones. Targets Heterocycl. Syst. 2014, 18, 113–140. [Google Scholar]
  7. Li, J.; Zhao, Y.-F.; Yuan, X.-Y.; Xu, J.-X.; Gong, P. Synthesis and Anticancer Activities of Novel 1,4-Disubstituted Phthalazines. Molecules 2006, 11, 574–582. [Google Scholar] [CrossRef] [Green Version]
  8. Zhai, X.; Li, J.; He, L.; Zheng, S.; Bin Zhang, Y.; Gong, P. Synthesis and in vitro cytotoxicity of novel 1,4-disubstituted phthalazines. Chin. Chem. Lett. 2008, 19, 29–32. [Google Scholar] [CrossRef]
  9. Carnovale, I.M.; Lolli, M.L.; Serra, S.C.; Mingo, A.F.; Napolitano, R.; Boi, V.; Guidolin, N.; Lattuada, L.; Tedoldi, F.; Baranyai, Z.; et al. Exploring the intramolecular catalysis of the proton exchange process to modulate the relaxivity of Gd(iii)-complexes of HP-DO3A-like ligands. Chem. Commun. 2018, 54, 10056–10059. [Google Scholar] [CrossRef]
  10. Chen, Y.-H.; Zhang, Y.-H.; Zhang, H.-J.; Liu, D.-Z.; Gu, M.; Li, J.-Y.; Wu, F.; Zhu, X.-Z.; Li, J.; Nan, F.-J. Design, Synthesis, and Biological Evaluation of Isoquinoline-1,3,4-trione Derivatives as Potent Caspase-3 Inhibitors. J. Med. Chem. 2006, 49, 1613–1623. [Google Scholar] [CrossRef]
  11. Ma, X.-Q.; Zhang, H.-J.; Zhang, Y.-H.; Chen, Y.-H.; Wu, F.; Du, J.-Q.; Yu, H.-P.; Zhou, Z.-L.; Li, J.-Y.; Nan, F.-J.; et al. Novel irreversible caspase-1 inhibitor attenuates the maturation of intracellular interleukin-1β. Biochem. Cell Boil. 2007, 85, 56–65. [Google Scholar] [CrossRef] [PubMed]
  12. Nan, F.-J.; Li, J.; Chen, L.-H.; Zhang, Y.-H.; Gu, M.; Zhang, H.-J. Isoquinoline-1,3,4-trione compounds, the method and the use thereof. US 20060135557, 2006. [Google Scholar]
  13. Zhang, Y.H.; Zhang, H.J.; Wu, F.; Chen, Y.H.; Ma, X.Q.; Du, J.Q.; Zhou, Z.L.; Li, J.Y.; Nan, F.J.; Li, J. Isoquinoline-1,3,4-trione and its derivatives attenuate beta-amyloid-induced apoptosis of neuronal cells. FEBS J. 2006, 273, 4842–4852. [Google Scholar] [CrossRef] [PubMed]
  14. Mitchell, G.; Clarke, E.D.; Ridley, S.M.; Greenhow, D.T.; Gillen, K.J.; Vohra, S.K.; Wardman, P. 1,3,4(2H)-isoquinolinetrione herbicides: Novel redox mediators of photosystem I. Pestic. Sci. 1995, 44, 49–58. [Google Scholar] [CrossRef]
  15. Mitchell, G.; Clarke, E.D.; Ridley, S.M.; Bartlett, D.W.; Gillen, K.J.; Vohra, S.K.; Greenhow, D.T.; Ormrod, J.C.; Wardman, P. 1,3,4(2H)-Isoquinolinetriones: Evaluation of amino-substituted derivatives as redox mediator herbicides. Pest Manag. Sci. 2000, 56, 120–126. [Google Scholar] [CrossRef]
  16. Mitchell, G.; Clarke, E.D.; Ridley, S.M.; Gillen, K.J.; Vohra, S.K.; Greenhow, D.T. Synthesis and characterisation of some 4-keto derivatives of 1,3,4(2H)-isoquinolinetrione redox mediator herbicides. Pest Manag. Sci. 2000, 56, 127–132. [Google Scholar] [CrossRef]
  17. Mazza, M.; Modena, T. Herbicidal activity of 2-substituted 1,3,4-(2H)-isoquinolinetriones. Il Farm. 1999, 54, 339–345. [Google Scholar] [CrossRef]
  18. Quattropani, A.; Dorbais, J.; Covini, D.; Pittet, P.-A.; Colovray, V.; Thomas, R.J.; Coxhead, R.; Halazy, S.; Scheer, A.; Missotten, M.; et al. Discovery and Development of a New Class of Potent, Selective, Orally Active Oxytocin Receptor Antagonists. J. Med. Chem. 2005, 48, 7882–7905. [Google Scholar] [CrossRef] [PubMed]
  19. Pollers-Wieërs, C.; Vekemans, J.; Toppet, S.; Hoornaert, G. The use of isoquinolinetriones in the synthesis of benzo[c]phenanthridine alkaloids. Tetrahedron 1981, 37, 4321–4326. [Google Scholar] [CrossRef]
  20. Wakchaure, P.B.; Easwar, S.; Puranik, V.G.; Argade, N.P. Facile air-oxidation of N-homopiperonyl-5,6-dimethoxyhomophthalimide: Simple and efficient access to nuevamine. Tetrahedron 2008, 64, 1786–1791. [Google Scholar] [CrossRef]
  21. Yoshifuji, S.; Arakawa, Y. Ruthenium Tetraoxide Oxidation of 3,4-Dihydroisoquinolin-1(2H)-ones: An Efficient Synthesis of Isoquinoline-1,3,4(2H)-triones. Chem. Pharm. Bull. 1989, 37, 3380–3381. [Google Scholar] [CrossRef]
  22. Ling, K.-Q.; Ye, J.-H.; Chen, X.-Y.; Ma, D.-J.; Xu, J.-H. On the reactions of 1,3-isoquinolinediones with singlet oxygen. Tetrahedron 1999, 55, 9185–9204. [Google Scholar] [CrossRef]
  23. Ling, K.-Q.; Ji, G.; Cai, H.; Xu, J.-H. Dye-Sensitized Photooxygenations of 1,3-Isoquinolinediones. Tetrahedron Lett. 1999, 39, 2381–2384. [Google Scholar]
  24. Vekemans, J.; Hoornaert, G. A new pathway to 1,3,4(2H)-isoquinolinetriones and substituted isoindolinones. Tetrahedron 1980, 36, 943–950. [Google Scholar] [CrossRef]
  25. Mahajan, S.; Sharma, B.; Kapoor, K.K. A solvent-free one step conversion of ketones to amides via Beckmann rearrangement catalysed by FeCl3·6H2O in presence of hydroxylamine hydrochloride. Tetrahedron Lett. 2015, 56, 1915–1918. [Google Scholar] [CrossRef]
  26. Yadav, J.; Reddy, B.S.; Reddy, U.S.; Praneeth, K. Azido-Schmidt reaction for the formation of amides, imides and lactams from ketones in the presence of FeCl3. Tetrahedron Lett. 2008, 49, 4742–4745. [Google Scholar] [CrossRef]
  27. Zhu, D.; Luo, W.-K.; Yang, L.; Ma, D.-Y. Iodine-catalyzed oxidative multiple C–H bond functionalization of isoquinolines with methylarenes: an efficient synthesis of isoquinoline-1,3,4(2H)-triones. Org. Biomol. Chem. 2017, 15, 7112–7116. [Google Scholar] [CrossRef] [PubMed]
  28. Al-Salahi, R.; Alswaidan, I.; Marzouk, M.; Iba, M. Cytotoxicity Evaluation of a New Set of 2-Aminobenzo[de]iso-quinoline-1,3-diones. Int. J. Mol. Sci. 2014, 15, 22483–22491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Tsou, H.R.; Otteng, M.; Tran, T.; Floyd, M., Jr.; Reich, M.; Birnberg, G.; Kutterer, K.; Ayral-Kaloustian, S.; Ravi, M.; Nilakantan, R.; et al. 4-(Phenylaminomethylene)isoquinoline-1,3(2H,4H)-diones as potent and selective inhibitors of the cyclin-dependent kinase 4 (CDK4). J. Med. Chem. 2008, 51, 3507–3525. [Google Scholar] [CrossRef] [PubMed]
  30. Billamboz, M.; Bailly, F.; Lion, C.; Touati, N.; Vezin, H.; Calmels, C.; Andréola, M.-L.; Christ, F.; Debyser, Z.; Cotelle, P. Magnesium Chelating 2-Hydroxyisoquinoline-1,3(2H,4H)-diones, as Inhibitors of HIV-1 Integrase and/or the HIV-1 Reverse Transcriptase Ribonuclease H Domain: Discovery of a Novel Selective Inhibitor of the Ribonuclease H Function. J. Med. Chem. 2011, 54, 1812–1824. [Google Scholar] [CrossRef]
  31. Ontoria, J.M.; Rydberg, E.H.; Di Marco, S.; Tomei, L.; Attenni, B.; Malancona, S.; Hernando, J.I.M.; Gennari, N.; Koch, U.; Narjes, F.; et al. Identification and Biological Evaluation of a Series of 1H-Benzo[de]isoquinoline-1,3(2H)-diones as Hepatitis C Virus NS5B Polymerase Inhibitors‡. J. Med. Chem. 2009, 52, 5217–5227. [Google Scholar] [CrossRef]
  32. Vernekar, S.K.V.; Liu, Z.; Nagy, E.; Miller, L.; Kirby, K.A.; Wilson, D.J.; Kankanala, J.; Sarafianos, S.T.; Parniak, M.A.; Wang, Z. Design, Synthesis, Biochemical, and Antiviral Evaluations of C6 Benzyl and C6 Biarylmethyl Substituted 2-Hydroxylisoquinoline-1,3-diones: Dual Inhibition against HIV Reverse Transcriptase-Associated RNase H and Polymerase with Antiviral Activities. J. Med. Chem. 2015, 58, 651–664. [Google Scholar] [CrossRef] [PubMed]
  33. Krawiecka, M.; Kossakowski, J.; Szymanek, K.; Kierzkowska, M.; Młynarczyk, G.; Kuran, B. Synthesis and antimicrobial activity of derivatives of 1H-benzo[de]isoquinoline-1,3(2H)-dione. Heterocycl. Commun. 2012, 18, 275–278. [Google Scholar]
  34. Staudinger, H.; Meyer, J. Über neue organische Phosphorverbindungen III. Phosphinmethylenderivate und Phosphinimine. Helv. Chim. Acta 1919, 635–646. [Google Scholar] [CrossRef]
  35. Di Mola, A.; Scorzelli, F.; Monaco, G.; Palombi, L.; Massa, A. Highly diastereo- and enantioselective organocatalytic synthesis of new heterocyclic hybrids isoindolinone-imidate and isoindolinone-phthalide. RSC Adv. 2016, 6, 60780–60786. [Google Scholar] [CrossRef]
  36. Thapa, P.; Corral, E.; Sardar, S.; Pierce, B.S.; Foss, F.W., Jr. Isoindolinone Synthesis: Selective Dioxane-Mediated Aerobic Oxidation of Isoindolines. J. Org. Chem. 2019, 84, 1025–1034. [Google Scholar] [CrossRef] [PubMed]
  37. Yang, Y.; Li, Y.; Cheng, C.; Yang, G.; Wan, S.; Zhang, J.; Mao, Y.; Zhao, Y.; Zhang, L.; Li, C.; et al. Reductant-Free Aerobic Hydroxylation of Isoquinoline-1,3(2H,4H)-dione Derivatives. J. Org. Chem. 2019, 84, 2316–2324. [Google Scholar] [CrossRef] [PubMed]
  38. Muchowski, J.M. One-step conversion of isatins to oxindoles and phthalonimides to homophthalimides. Can. J. Chem. 1969, 47, 857–859. [Google Scholar] [CrossRef]
  39. Zeneca, A. Isoquinolinetrione derivatives as herbicides. Pat. appln PCT/GB/94/01094, 1994. [Google Scholar]
Sample Availability: Samples of the compounds are available from the authors.
Scheme 1. This work and literature methodologies for the synthesis of 2,3-dihydroisoquinoline-1,3,4-trione.
Scheme 1. This work and literature methodologies for the synthesis of 2,3-dihydroisoquinoline-1,3,4-trione.
Molecules 24 02177 sch001
Figure 1. 1,3,4(2H)-Isoquinolinetrione derivatives of biological interest.
Figure 1. 1,3,4(2H)-Isoquinolinetrione derivatives of biological interest.
Molecules 24 02177 g001
Figure 2. ORTEP drawing and atom numbering scheme of compound 3a. Hydrogen atoms have been omitted for clarity. Ellipsoids are drawn at 30% probability level. CCDC 1915668 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, Cambridge, UK).
Figure 2. ORTEP drawing and atom numbering scheme of compound 3a. Hydrogen atoms have been omitted for clarity. Ellipsoids are drawn at 30% probability level. CCDC 1915668 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, Cambridge, UK).
Molecules 24 02177 g002
Scheme 2. Synthesis of NH-unsubstitued 2,3-dihydroisoquinoline-1,3,4-trione.
Scheme 2. Synthesis of NH-unsubstitued 2,3-dihydroisoquinoline-1,3,4-trione.
Molecules 24 02177 sch002
Scheme 3. Proposed one-pot mechanism for the developed process.
Scheme 3. Proposed one-pot mechanism for the developed process.
Molecules 24 02177 sch003
Scheme 4. Proposed mechanism for air oxidation to afford 2,3-dihydroisoquinoline-1,3,4-triones.
Scheme 4. Proposed mechanism for air oxidation to afford 2,3-dihydroisoquinoline-1,3,4-triones.
Molecules 24 02177 sch004
Table 1. Preliminary reaction conditions screening.
Table 1. Preliminary reaction conditions screening.
Molecules 24 02177 i001
EntryBase (1 eq)T (°C)Time (h)Yield 3a a
1-r.t.3--
2-r.t.18--
3DIPEAr.t1818% b
4DIPEA50353%
5DIPEA501871%
6cDIPEA501810%
7Et3N501850%
8NaHCO3r.t1826%
9NaHCO350341%
10K2CO350440%
11K2CO35018Dec.
a Yields refer to chromatographically pure compounds. b 80% of 1 was recovered. c Reaction was performed under nitrogen.
Table 2. Substrate scope.
Table 2. Substrate scope.
Molecules 24 02177 i002
EntryR3Yield % a
1-CH2Ph(3a)71%
24-Cl-CH2Ph(3b)73%
34-MeO-CH2Ph(3c)75%
42-F-CH2Ph(3d)69%
53-NO2-CH2Ph(3e)69%
6-CH3(3f)72%
7-CH2CH3(3g)70%
8
9
-(CH2)3CH3
-CH2CH=CH2
(3h)
(3i)
73%
72%
10Ph --
a Yields refer to chromatographically pure compounds.

Share and Cite

MDPI and ACS Style

Di Mola, A.; Tedesco, C.; Massa, A. Metal-Free Air Oxidation in a Convenient Cascade Approach for the Access to Isoquinoline-1,3,4(2H)-triones. Molecules 2019, 24, 2177. https://doi.org/10.3390/molecules24112177

AMA Style

Di Mola A, Tedesco C, Massa A. Metal-Free Air Oxidation in a Convenient Cascade Approach for the Access to Isoquinoline-1,3,4(2H)-triones. Molecules. 2019; 24(11):2177. https://doi.org/10.3390/molecules24112177

Chicago/Turabian Style

Di Mola, Antonia, Consiglia Tedesco, and Antonio Massa. 2019. "Metal-Free Air Oxidation in a Convenient Cascade Approach for the Access to Isoquinoline-1,3,4(2H)-triones" Molecules 24, no. 11: 2177. https://doi.org/10.3390/molecules24112177

Article Metrics

Back to TopTop