Next Article in Journal
Phenanthroline-Mediated Photoelectrical Enhancement in Calix[4]arene-Functionalized Titanium-Oxo Clusters
Previous Article in Journal
Impact of Anchoring Groups on the Photocatalytic Performance of Iridium(III) Complexes and Their Toxicological Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of N-Aryl-Substituted Pyrrolidines by Successive Reductive Amination of Diketones via Transfer Hydrogenation

1
School of Pharmaceutical Sciences, Gannan Medical University, Ganzhou 341000, China
2
College of Chemistry and Environmental Engineering, Shaoguan University, Shaoguan 512005, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(11), 2565; https://doi.org/10.3390/molecules29112565
Submission received: 29 April 2024 / Revised: 19 May 2024 / Accepted: 21 May 2024 / Published: 30 May 2024
(This article belongs to the Section Organic Chemistry)

Abstract

:
N-aryl-substituted pyrrolidines are important moieties widely found in bioactive substances and drugs. Herein, we present a practical reductive amination of diketones with anilines for the synthesis of N-aryl-substituted pyrrolidines in good to excellent yields. In this process, the N-aryl-substituted pyrrolidines were furnished via successive reductive amination of diketones via iridium-catalyzed transfer hydrogenation. The scale-up performance, water as a solvent, simple operation, as well as derivation of drug molecules showcased the potential application in organic synthesis.

1. Introduction

N-aryl-substituted pyrrolidines are important structural fragments that widely exist in bioactive substances (Scheme 1) [1,2,3,4,5,6,7]. As such, a variety of methods for constructing these compounds have been designed and developed for this purpose [8]. The amination of aromatic amines with dihalogenated compounds [9,10,11,12] or diols [13,14,15] was perceived as a prevalent classical method for pyrrolidine synthesis. Additionally, reductive amination is also one of the general strategies for building N-aryl-substituted pyrrolidines [16,17,18,19,20,21,22,23]. Additionally, the hydrogenation of N-heteroaromatics was employed as well for the construction of N-aryl-substituted pyrrolidines [24,25,26]. It is worth noting that other methods, including cross-coupling of aryl halides [27], reduction in tertiary amides [28] and lactams [29,30], intramolecular C-N coupling [31] or C-N amination [32], reactions of cyclic ether compounds with primary arylamines [33,34,35,36,37] or arylhydrazines [38], Mitsunobu cyclodehydration reaction [39], Prins cyclization [40,41], aminomercuration demercuration [42], as well as intramolecular carbonyl olefination of amides [43], were established to furnish N-aryl-substituted pyrrolidines. Furthermore, chiral-substituted pyrrolidines could be delivered via copper-catalyzed asymmetric 1,3-diploar cycloaddition [44].
Despite the great progress achieved in the construction of N-aryl-substituted pyrrolidines, challenges still exist in terms of selectivity, substrate versatility, reaction conditions, as well as catalytic efficiency [45,46]. Among the general strategies for pyrrolidine synthesis, reductive amination is undoubtedly the most efficient approach. In this process, the reaction of carbonyl compounds with amines took place first to afford C=N intermediates, which was then followed by the reduction process to form C-N bonds, in which water is the only by-product. Nevertheless, an excess of expensive hydrides or silanes was generally required as a reductant in these traditional processes [16,17,18,19,20,21,22,23]. Therefore, a cheap and efficient catalytic system is still in demand for pyrrolidine synthesis.
Metal-catalyzed transfer hydrogenation employed as a supplement to conventional hydrogenation drew much attention and was extensively utilized in organic synthesis [47]. In this context, hydrogen donors such as formic acid were widely utilized in differential transfer hydrogenation reactions owing to their potential hydrogen storage, low toxicity, as well as ease of handling [48]. We have great interest in transfer hydrogenation and conducted a series of relative studies using iridium complexes as catalysts [49], by which reduction in unsaturated compounds [50,51], silane oxidation [52], N-allylic alkylation [53], reductive amination [54], reductive amination [55], as well as reductive etherification [56] were reported. Recently, we established a one-pot procedure for synthesizing amines via reduction-reductive amination (Scheme 2a) [57]. Based on previous work, we envisioned that this approach could also be employed for the construction of N-aryl-substituted pyrrolidines via successive reductive amination. Herein, we describe an iridium-catalyzed successive reductive amination of diketones with anilines, offering N-aryl-substituted pyrrolidines in good to excellent yields under mild conditions (Scheme 2b). This protocol provides a new route for azacycle synthesis.

2. Results and Discussion

The development of this iridium-catalyzed reductive amination reaction began with the model reaction of 2,5-hexanedione (1a) with aniline (2a) (Table 1). Catalyst screening indicated similar yields of mixed products of N-phenyl-substituted pyrrolidine 3a1 and N-phenyl-substituted pyrrole 3a1′ were formed simultaneously (Table 1, entries 1–6). To afford the appropriate conditions for N-phenyl-substituted pyrrolidine synthesis, further reaction conditions were screened. Solvent optimization (Table 1, entries 7–13) showed that water was beneficial to the formation of the desired pyrrolidine 3a1 in 80% yield, in which the pyrrole product 3a1′ was inhibited (Table 1, entry 12). Notably, the increasing loading of HCO2H is helpful for the transformation of the desired pyrrolidine product 3a1 (Table 1, entries 14–17). For instance, a 92% isolated yield of N-phenyl-substituted pyrrolidine 3a1 was generated in 71:29 dr by increasing the dosage of HCO2H to 30.0 equivalent (Table 1, entry 17). Similar excellent yields of pyrrolidine 3a1 were afforded by increasing the reaction temperature to 100 °C, even shortening the reaction time (Table 1, entries 18 and 19). In comparison, only 60% yield of 3a1 in 50:50 dr was obtained when the reaction was performed at room temperature (Table 1, entry 20). Control experiments showed both the catalyst and formic acid were essential for this transformation (Table 1, entries 21 and 22).
With the above optimized conditions in hand, differentially substituted aromatic amines were explored for the synthesis of N-aryl-substituted pyrrolidines using 2,5-hexanedione (1a) as substrate. As showcased in Scheme 3, para-substituted aromatic amines with different electronic effects and positions were well tolerated in this system, affording the desired products of 3a23a9 in moderate to excellent yields and stereoselectivities. Notably, reductive amination of 2,5-hexanedione (1a) with 2-naphthylamine (2m), 5,6,7,8-tetrahydro-naphthalen-2-amine (2n), 2,3-dihydro-1H-inden-5-amine (2o), 1-methylindolin-6-amine (2p), as well as 4-amino-N-phenylbenzamide (2q) also produced the corresponding N-aryl-substituted pyrrolidines 3a133a17 in moderate yields and stereoselectivities. Interestingly, similar yield and stereoselectivity of the procaine-derived pyrrolidine 3a18 were afforded as well by using procaine as a nucleophile. However, both pyrrolidine (3a19) and pyrrole (3a19′) products were formed simultaneously in similar yield with 5-bromonaphthalen-1-amine (2s) as substrate.
To further explore the practicability of this catalytic system, the scope of diketones for this reductive amination with different aromatic amines was screened as well. As shown in Scheme 4, similar moderate yields but excellent stereoselectivities (>99:1 dr) of N-aryl-substituted pyrrolidines (3b1, 3b2~3b4) were afforded via reductive amination of 1-phenylhexane-1,4-dione (1b) with para-substituted aromatic amines, in which the configuration of 3b4 was determined by X-ray crystallography [58]. However, decreased stereoselectivities of the desired product (3b5, 3b6) were observed using ortho- and para-substituted aromatic amines as nucleophiles and 1-phenylhexane-1,4-dione (1b) as substrate. Obviously, employing more steric hinderance of octane-3,6-dione (1c) as substrate resulted in similar moderate yields and stereoselectivities (3c1 and 3c2). Surprisingly, the formation of pyrrole as a major product (3c3′) was observed by using 4-iodoaniline (2y) as a nucleophile, delivering a reductive amination product of 3c3 in lower yield.
To further illustrate the utility of this iridium-catalyzed reductive amination, the model reaction was subjected to a gram-scale performance under standard conditions (Scheme 5). As expected, 10.0 mmol of the compound 1a was successfully converted into 1.61 g of the corresponding product 3a1 in a 92% yield.
To shed light on the process of this Ir-catalyzed reductive amination, control experiments were conducted to explore more details of this transformation. Firstly, to probe whether the pyrrole is involved as an intermediate in this reductive amination process, a control experiment using 3a1′ as substrate was performed under standard conditions (Scheme 6a). However, no expected pyrrolidine product 3a1 was observed, strongly suggesting that the pyrrolidine product was not produced via the transfer hydrogenation of pyrrolidine.
Based on the experimental results and related studies [16,55,57], a possible process of this Ir-catalyzed successive reductive amination was proposed in Scheme 7. Comparative pathways took place in the presence of diketones and amines under these standard conditions. On the one hand, the Paalknorr condensation took place to deliver the pyrrole product of 3′ (Path A) [59]. On the other hand, the imidone intermediate 6 was formed firstly via condensation, which was reduced to afford the amino ketone 7 via transfer hydrogenation under the action of iridium catalysts (Path B). Using amino ketone 7 as material, a more similar condensation and transfer hydrogenation were undergone to produce the pyrrolidine product 3. In this process, we envisaged that the stable five-membered ring intermediate 9 was crucial to deliver the corresponding pyrrolidine product, which was supported by the formation of amino alcohol 5a1, rather than the four-membered ring product using pentane-2,4-dione (4a) as substrate (Scheme 6b).

3. Experimental Section

All the reactions were carried out in oven-dried Schlenk tubes. All the reagents and anhydrous solvents were purchased from commercial sources and used without further purification. Silica gel (100~200 mesh) bought from commercial sources was used for column chromatography. Purified hexane or a mixture of petroleum ether and ethyl acetate was used as a gradient eluent for column chromatography. 1H and 13C NMR spectra were recorded using a Bruker DRX-400 spectrometer (Bruker, Ettlingen, Germany) (400 MHz for 1H; 101 MHz for 13C). The chemical shifts are referenced to the resonances of the residual protons in the deuterated solvents. The abbreviations [s = singlet, d = doublet, t = triplet, m = multiplet, br = broad coupling constants are given in Hertz (Hz)] were used to designate the chemical shift multiplicities. All NMR quantitative analyses were performed with dimethyl terephthalate as the internal standard. High-resolution mass spectra (HRMS) were recorded by an LCMS-IT-TOF mass spectrometer. Melting points were obtained on the WRR melting point apparatus without correction. Fourier transform infrared spectra (FT-IR) were recorded on a Thermofisher Nicolet iS50 (Thermo, Waltham, MA, USA) spectrometer.

3.1. General Procedure for Construction of N-Aryl-Substituted Pyrrolidines

To a 25.0 mL dried Schlenk tube, add catalyst TC-2 [60] (1.0 mol%), 1 (0.5 mmol, 1.0 equiv.), 2 (0.6 mmol, 1.1 equiv.), solvent (2.0 mL), and hydrogen donor HCO2H (30.0 equiv.), which was stirred under air at 80 °C for 12 h. After the completion of the reaction, the mixture was dissolved in ethyl acetate and washed with saturated salt water 2~3 times, and then the organic fraction was dried over anhydrous Na2SO4. The residue was separated and refined by chromatography on silica gel with hexane or a mixture of petroleum ether and ethyl acetate as the eluent to afford product 3.

3.2. Large-Scale Synthesis of 3a1

To a 100.0-mL dried round bottom schleck, catalyst TC-2 (1.0 mol%, 57.92 mg), 1a (10.0 mmol, 1.17 mL), 2a (12.0 mmol, 1.09 mL), solvent (30.0 mL), and hydrogen donor HCO2H (300.0 mmol, 1.13 mL) were added, which were stirred under air at 80 °C for 12 h. The mixture was diluted with EtOAc (20.0 mL) and then quenched with saturated salt water (30.0 mL). The organic layer was dried over anhydrous Na2SO4. The residue was purified by chromatography on silica gel with hexane to afford the product 3a1.

4. Conclusions

Overall, we have developed a practical successive reductive amination of diketones with anilines for the construction of N-aryl-substituted pyrrolidines in good to excellent yields and stereoselectivities. This method features scale-up performance, water as a solvent, simple operation, as well as the derivation of drug molecules. Mechanistic studies indicated competitive pathways of Paal-knorr condensation reaction and Ir-catalyzed transfer hydrogenation to produce pyrrole and pyrrolidine products, respectively. Further efforts into the mechanistic details as well as explorations of the medical applications of the N-aryl-substituted pyrrolidine products are currently underway in our laboratory.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29112565/s1, Figure S1: X-ray crystal structures of compound 3b4; Table S1: Crystal data and structure refinement for 3b4.

Author Contributions

Conceptualization, R.L.; Methodology, J.T. and L.L.; Resources, R.L.; Data curation, L.O.; Writing—original draft, J.L.; Writing—review & editing, R.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (22161004), the Fundamental Research Funds for Gannan Medical University (QD202019, QD202106, TD2021YX05) and Shaoguan University (408/9900064703) for financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. O’Hagan, D. Pyrrole, pyrrolidine pyridine, piperidine, azepine and tropane alkaloids. Nat. Prod. Rep. 1997, 14, 637–651. [Google Scholar] [CrossRef]
  2. Balaban, A.T.; Oniciu, D.C.; Katritzky, A.R. Aromaticity as a Cornerstone of Heterocyclic Chemistry. Chem. Rev. 2004, 104, 2777–2812. [Google Scholar] [CrossRef] [PubMed]
  3. La Regina, G.; Silvestri, R.; Artico, M.; Lavecchia, A.; Novellino, E.; Befani, O.; Turini, P.; Agostinelli, E. New Pyrrole inhibitors of monoamine oxidase: Synthesis, biological evaluation, and structural determinants of MAO-A and MAO-B selectivity. J. Med. Chem. 2007, 50, 922–931. [Google Scholar] [CrossRef] [PubMed]
  4. Asahina, Y.; Takei, M.; Kimura, T.; Fukuda, Y. Synthesis and antibacterial activity of novel pyrido [1, 2, 3-de][1, 4] benzoxazine-6-carboxylic acid derivatives carrying the 3-cyclopropylaminomethyl-4-substituted-1-pyrrolidinyl group as a C-10 substituent. J. Med. Chem. 2008, 51, 3238–3249. [Google Scholar] [CrossRef] [PubMed]
  5. Srinivasan, V.; Panneerselvam, M.; Pavithra, N.; Anandan, S.; Sundaravel, K.; Jaccob, M.; Kathiravan, A. A combined experimental and computational characterization of D–π–A dyes containing heterocyclic electron donors. J. Photochem. Photobiol. A Chem. 2017, 332, 453–464. [Google Scholar] [CrossRef]
  6. Reddy, B.V.S.; Nair, P.N.; Antony, A.; Lalli, C.; Grée, R. The aza-Prins reaction in the synthesis of natural products and analogues. Eur. J. Org. Chem. 2017, 2017, 1805–1819. [Google Scholar] [CrossRef]
  7. Verma, P.; Richter, J.M.; Chekshin, N.; Qiao, J.X.; Yu, J.-Q. Iridium (I)-catalyzed α-C (sp3)-H Alkylation of Saturated Azacycles. J. Am. Chem. Soc. 2020, 142, 5117–5125. [Google Scholar] [CrossRef]
  8. Popovici-Muller, J.; Lemieux, R.M.; Artin, E.; Saunders, J.O.; Salituro, F.G.; Travins, J.; Cianchetta, G.; Cai, Z.; Zhou, D.; Cui, D.; et al. Discovery of AG-120 (Ivosidenib): A First-in-Class Mutant IDH1 Inhibitor for the Treatment of IDH1 Mutant Cancers. ACS Med. Chem. Lett. 2018, 9, 300–305. [Google Scholar] [CrossRef]
  9. Ju, Y.; Varma, R.S. An Efficient and Simple Aqueous N-Heterocyclization of Aniline Derivatives: Microwave-Assisted Synthesis of N-Aryl Azacycloalkanes. Org. Lett. 2005, 7, 2409–2411. [Google Scholar] [CrossRef]
  10. Ju, Y.; Varma, R.S. Aqueous N-Heterocyclization of Primary Amines and Hydrazines with Dihalides: Microwave-Assisted Syntheses of N-Azacycloalkanes, Isoindole, Pyrazole, Pyrazolidine, and Phthalazine Derivatives. J. Org. Chem. 2006, 71, 135–141. [Google Scholar] [CrossRef]
  11. Barnard, T.M.; Vanier, G.S.; Collins, M.J. Scale-up of the green synthesis of azacycloalkanes and isoindolines under microwave irradiation. Org. Process Res. Dev. 2006, 10, 1233–1237. [Google Scholar] [CrossRef]
  12. Gonsalves, O.S.; Ambre, J.P.; Nemade, P.R. Improving the yield of graphene oxide-catalysed N-heterocyclization of amines through fed batch mode. New J. Chem. 2022, 46, 17410–17420. [Google Scholar] [CrossRef]
  13. Cui, X.; Dai, X.; Deng, Y.; Shi, F. Development of a General Non-Noble Metal Catalyst for the Benign Amination of Alcohols with Amines and Ammonia. Chem. Eur. J. 2013, 19, 3665–3675. [Google Scholar] [CrossRef] [PubMed]
  14. Yang, P.; Zhang, C.; Gao, W.-C.; Ma, Y.; Wang, X.; Zhang, L.; Yue, J.; Tang, B. Nickel-catalyzed borrowing hydrogen annulations: Access to diversified N-heterocycles. Chem. Commun. 2019, 55, 7844–7847. [Google Scholar] [CrossRef] [PubMed]
  15. Chamberlain, A.E.R.; Paterson, K.J.; Armstrong, R.J.; Twin, H.C.; Donohoe, T.J. A hydrogen borrowing annulation strategy for the stereocontrolled synthesis of saturated aza-heterocycles. Chem. Commun. 2020, 56, 3563–3566. [Google Scholar] [CrossRef]
  16. Boga, C.; Manescalchi, F.; Savoia, D. Diastereoselective synthesis of 2,5-dimethylpyrrolidines and 2,6-dimethylpiperidines by reductive amination of 2,5-hexanedione and 2,6-heptanedione with hydride reagents. Tetrahedron 1994, 50, 4709–4722. [Google Scholar] [CrossRef]
  17. Xue, Z.; Yu, D.; Zhao, X.; Mu, T. Upgrading of levulinic acid into diverse N-containing functional chemicals. Green Chem. 2019, 21, 5449–5468. [Google Scholar] [CrossRef]
  18. Wang, S.; Huang, H.; Bruneau, C.; Fischmeister, C. Selective and Efficient Iridium Catalyst for the Reductive Amination of Levulinic Acid into Pyrrolidones. ChemSusChem 2017, 10, 4150–4154. [Google Scholar] [CrossRef]
  19. Wu, C.; Luo, X.; Zhang, H.; Liu, X.; Ji, G.; Liu, Z.; Liu, Z. Reductive amination/cyclization of levulinic acid to pyrrolidones versus pyrrolidines by switching the catalyst from AlCl3 to RuCl3 under mild conditions. Green Chem. 2017, 19, 3525–3529. [Google Scholar] [CrossRef]
  20. Wu, C.; Zhang, H.; Yu, B.; Chen, Y.; Ke, Z.; Guo, S.; Liu, Z. Lactate-based ionic liquid catalyzed reductive amination/cyclization of keto acids under mild conditions: A metal-free route to synthesize lactams. ACS Catal. 2017, 7, 7772–7776. [Google Scholar] [CrossRef]
  21. Muzzio, M.; Yu, C.; Lin, H.; Yom, T.; Boga, D.A.; Xi, Z.; Li, N.; Yin, Z.; Li, J.; Dunn, J.A. Reductive amination of ethyl levulinate to pyrrolidones over AuPd nanoparticles at ambient hydrogen pressure. Green Chem. 2019, 21, 1895–1899. [Google Scholar] [CrossRef]
  22. Xie, C.; Song, J.; Wu, H.; Hu, Y.; Liu, H.; Zhang, Z.; Zhang, P.; Chen, B.; Han, B. Ambient Reductive Amination of Levulinic Acid to Pyrrolidones over Pt Nanocatalysts on Porous TiO2 Nanosheets. J. Am. Chem. Soc. 2019, 141, 4002–4009. [Google Scholar] [CrossRef]
  23. Irrgang, T.; Kempe, R. Transition-metal-catalyzed reductive amination employing hydrogen. Chem. Rev. 2020, 120, 9583–9674. [Google Scholar] [CrossRef]
  24. Duan, Y.; Li, L.; Chen, M.-W.; Yu, C.-B.; Fan, H.-J.; Zhou, Y.-G. Homogenous Pd-catalyzed ASYMMETRIC HYDROGENATION of unprotected indoles: Scope and mechanistic studies. J. Am. Chem. Soc. 2014, 136, 7688–7700. [Google Scholar] [CrossRef]
  25. Hu, X.-N.; Shen, T.-L.; Cai, D.-C.; Zheng, J.-F.; Huang, P.-Q. The iridium-catalysed reductive coupling reaction of tertiary lactams/amides with isocyanoacetates. Org. Chem. Front. 2018, 5, 2051–2056. [Google Scholar] [CrossRef]
  26. Menuel, S.; Bertaut, E.; Monflier, E.; Hapiot, F. Cyclodextrin-based PNN supramolecular assemblies: A new class of pincer-type ligands for aqueous organometallic catalysis. Dalton Trans. 2015, 44, 13504–13512. [Google Scholar] [CrossRef]
  27. Ruiz-Castillo, P.; Buchwald, S.L. Applications of palladium-catalyzed C–N cross-coupling reactions. Chem. Rev. 2016, 116, 12564–12649. [Google Scholar] [CrossRef]
  28. Snelling, R.A.; Amberchan, G.; Resendez, A.; Murphy, C.L.; Porter, L.; Singaram, B. Diisobutylaluminum borohydride: An efficient reagent for the reduction of tertiary amides to the corresponding amines under ambient conditions. Tetrahedron Lett. 2017, 58, 4073–4077. [Google Scholar] [CrossRef]
  29. Collins, C.J.; Lanz, M.; Singaram, B. Facile reduction of tertiary lactams to cyclic amines with 9-borabicyclo [3.3. 1] nonane (9-BBN). Tetrahedron Lett. 1999, 40, 3673–3676. [Google Scholar] [CrossRef]
  30. Flaniken, J.M.; Collins, C.J.; Lanz, M.; Singaram, B. Aminoborohydrides. 11. Facile reduction of N-alkyl lactams to the corresponding amines using lithium aminoborohydrides. Org. Lett. 1999, 1, 799–801. [Google Scholar] [CrossRef]
  31. Jeffrey, J.L.; Bartlett, E.S.; Sarpong, R. Intramolecular C(sp3)-N Coupling by Oxidation of Benzylic C, N-Dianions. Angew. Chem. Int. Ed. 2013, 52, 2194–2197. [Google Scholar] [CrossRef]
  32. Betz, K.N.; Chiappini, N.D.; Du Bois, J. Intermolecular sp3-C–H amination for the synthesis of saturated azacycles. Org. Lett. 2020, 22, 1687–1691. [Google Scholar] [CrossRef]
  33. Korbad, B.L.; Lee, S.H. Synthesis of N-aryl substituted, five-and six-membered azacycles using aluminum-amide complexes. Chem. Commun. 2014, 50, 8985–8988. [Google Scholar] [CrossRef]
  34. Amara, Z.; Streng, E.S.; Skilton, R.A.; Jin, J.; George, M.W.; Poliakoff, M. Automated Serendipity with Self-Optimizing Continuous-Flow Reactors. Eur. J. Org. Chem. 2015, 2015, 6141–6145. [Google Scholar] [CrossRef]
  35. Zhang, Z.; Miao, C.; Xia, C.; Sun, W. Synergistic acid-catalyzed synthesis of N-aryl-substituted azacycles from anilines and cyclic ethers. Org. Lett. 2016, 18, 1522–1525. [Google Scholar] [CrossRef]
  36. Sun, Z.; Hu, S.; Huo, Y.; Wang, Z. Titanium tetrachloride-mediated synthesis of N-aryl-substituted azacycles from cyclic ethers. RSC Adv. 2017, 7, 4363–4367. [Google Scholar] [CrossRef]
  37. La, M.T.; Kang, S.; Kim, H.-K. Metal-Free Synthesis of N-Aryl-Substituted Azacycles from Cyclic Ethers Using POCl3. J. Org. Chem. 2019, 84, 6689–6696. [Google Scholar] [CrossRef]
  38. Tran, V.H.; Hong, W.P.; Kim, H.K. Facile titanium (IV) chloride and TBD-mediated synthesis of N-aryl-substituted azacycles from arylhydrazines. Bull. Korean Chem. Soc. 2022, 43, 777–783. [Google Scholar] [CrossRef]
  39. Gill, D.M.; Iveson, M.; Collins, I.; Jones, A.M. A Mitsunobu reaction to functionalized cyclic and bicyclic N-arylamines. Tetrahedron Lett. 2018, 59, 238–242. [Google Scholar] [CrossRef]
  40. Miranda, P.O.; Carballo, R.M.; Martín, V.S.; Padrón, J.I. A new catalytic Prins cyclization leading to Oxa- and Azacycles. Org. Lett. 2009, 11, 357–360. [Google Scholar] [CrossRef]
  41. Liu, G.-Q.; Cui, B.; Xu, R.; Li, Y.-M. Preparation of trans-2-Substituted-4-halopiperidines and cis-2-Substituted-4-halotetrahydropyrans via AlCl3-Catalyzed Prins Reaction. J. Org. Chem. 2016, 81, 5144–5161. [Google Scholar] [CrossRef] [PubMed]
  42. Barluenga, J.; Nájera, C.; Yus, M. One-Pot synthesis of substituted pyrrolidines via aminomercuration-demercuration of 1, 4- and 1, 5-Hexadiene. J. Heterocyclic Chem. 1981, 18, 1297–1299. [Google Scholar] [CrossRef]
  43. Takeda, T.; Saito, J.; Tsubouchi, A. Formation of pyrrolidines by the titanocene (II)-promoted intramolecular reaction of N-[3, 3-bis (phenylthio) propyl] anilides. Tetrahedron Lett. 2003, 44, 5571–5574. [Google Scholar] [CrossRef]
  44. Qiu, H.; Du, Z.; Zhao, Y.; Yuan, S.; Xi, S.; Zhou, T.; Yang, J.; Zhang, C.; Xiong, Y.; Xia, Y.; et al. Enantioselective Synthesis and Biological Evaluation of Pyrrolidines Bearing Quaternary Stereogenic Centers. J. Med. Chem. 2023, 66, 9866–9880. [Google Scholar] [CrossRef] [PubMed]
  45. Fujita, K.-I.; Fujii, T.; Yamaguchi, R. Cp*Ir Complex-Catalyzed N-Heterocyclization of Primary Amines with Diols: A New Catalytic System for Environmentally Benign Synthesis of Cyclic Amines. Org. Lett. 2004, 6, 3525–3528. [Google Scholar] [CrossRef] [PubMed]
  46. He, L.; Lou, X.B.; Ni, J.; Liu, Y.M.; Cao, Y.; He, H.Y.; Fan, K.N. Efficient and Clean Gold-Catalyzed One-Pot Selective N-Alkylation of Amines with Alcohols. Chem. Eur. J. 2010, 16, 13965–13969. [Google Scholar] [CrossRef] [PubMed]
  47. Wu, X.; Wang, C.; Xiao, J. Transfer Hydrogenation in Water. Chem. Rec. 2016, 16, 2772–2786. [Google Scholar] [CrossRef] [PubMed]
  48. Nie, R.; Tao, Y.; Nie, Y.; Lu, T.; Wang, J.; Zhang, Y.; Lu, X.; Xu, C.C. Recent advances in catalytic transfer hydrogenation with formic acid over heterogeneous transition metal catalysts. ACS Catal. 2021, 11, 1071–1095. [Google Scholar] [CrossRef]
  49. Wei, Y.; Liang, Y.; Luo, R.; Ouyang, L. Recent advances of Cp*Ir complexes for transfer hydrogenation: Focus on formic acid/formate as hydrogen donors. Org. Biomol. Chem. 2023, 21, 7484–7497. [Google Scholar] [CrossRef]
  50. Xia, Y.; Wang, S.; Miao, R.; Liao, J.; Ouyang, L.; Luo, R. Synthesis of N-alkoxy amines and hydroxylamines via the iridium-catalyzed transfer hydrogenation of oximes. Org. Biomol. Chem. 2022, 20, 6394–6399. [Google Scholar] [CrossRef]
  51. Wen, H.; Luo, N.; Zhu, Q.; Luo, R. Amide iridium complexes as catalysts for transfer hydrogenation reduction of N-sulfonylimine. J. Org. Chem. 2021, 86, 3850–3859. [Google Scholar] [CrossRef] [PubMed]
  52. Luo, N.; Liao, J.; Ouyang, L.; Wen, H.; Zhong, Y.; Liu, J.; Tang, W.; Luo, R. Highly selective hydroxylation and alkoxylation of silanes: One-pot silane oxidation and reduction of aldehydes/ketones. Organometallics 2020, 39, 165–171. [Google Scholar] [CrossRef]
  53. Luo, N.; Zhong, Y.; Shui, H.; Luo, R. pH-Mediated Selective Synthesis of N-Allylic Alkylation or N-Alkylation Amines with Allylic Alcohols via an Iridium Catalyst in Water. J. Org. Chem. 2021, 86, 15509–15521. [Google Scholar] [CrossRef] [PubMed]
  54. Ouyang, L.; Miao, R.; Yang, Z.; Luo, R. Iridium-catalyzed reductive amination of carboxylic acids. J. Catal. 2023, 418, 283–289. [Google Scholar] [CrossRef]
  55. Ouyang, L.; Xia, Y.; Liao, J.; Luo, R. One-Pot Transfer Hydrogenation Reductive Amination of Aldehydes and Ketones by Iridium Complexes “on Water”. Eur. J. Org. Chem. 2020, 2020, 6387–6391. [Google Scholar] [CrossRef]
  56. Ouyang, L.; Xia, Y.; Miao, R.; Liao, J.; Luo, R. Iridium-catalyzed reductive etherification of α, β-unsaturated ketones and aldehydes with alcohols. Org. Biomol. Chem. 2022, 20, 2621–2625. [Google Scholar] [CrossRef]
  57. Luo, R.; Tong, J.; Ouyang, L.; Liu, L.; Liao, J. One-pot reductive amination of carbonyl compounds and nitro compounds via Ir-catalyzed transfer hydrogenation. RSC Adv. 2023, 13, 29607–29612. [Google Scholar] [CrossRef]
  58. CCDC Number: 2296401. Available online: https://www.ccdc.cam.ac.uk/structures/Search?access=referee&ccdc=2296401&Author=Renshi+Luo (accessed on 28 April 2024).
  59. Cho, H.; Madden, R.; Nisanci, B.; Török, B. The Paal-Knorr reaction revisited. A catalyst and solvent-free synthesis of underivatized and N-substituted pyrroles. Green Chem. 2015, 17, 1088–1099. [Google Scholar] [CrossRef]
  60. Yang, Z.; Zhu, Z.; Luo, R.; Qiu, X.; Liu, J.-T.; Yang, J.-K.; Tang, W. Iridium-catalyzed highly efficient chemoselective reduction of aldehydes in water using formic acid as the hydrogen source. Green Chem. 2017, 19, 3296–3301. [Google Scholar] [CrossRef]
Scheme 1. Bioactive molecules with N-aryl-substituted pyrrolidine moiety.
Scheme 1. Bioactive molecules with N-aryl-substituted pyrrolidine moiety.
Molecules 29 02565 sch001
Scheme 2. Iridium-catalyzed reductive amination for C-N bond formation.
Scheme 2. Iridium-catalyzed reductive amination for C-N bond formation.
Molecules 29 02565 sch002
Scheme 3. Synthesis of N-aryl-substituted pyrrolidines with various amines a. [a] Reaction conditions: 1 (0.5 mmol, 1.0 equiv.), 2 (0.6 mmol, 1.1 equiv.), solvent (2.0 mL), catalyst (1.0 mol%), HCO2H (30.0 equiv.), under air, at 80 °C, for 12 h. [b] Isolated yield. [c] Determined by 1H NMR analysis.
Scheme 3. Synthesis of N-aryl-substituted pyrrolidines with various amines a. [a] Reaction conditions: 1 (0.5 mmol, 1.0 equiv.), 2 (0.6 mmol, 1.1 equiv.), solvent (2.0 mL), catalyst (1.0 mol%), HCO2H (30.0 equiv.), under air, at 80 °C, for 12 h. [b] Isolated yield. [c] Determined by 1H NMR analysis.
Molecules 29 02565 sch003
Scheme 4. Ir-catalyzed reductive amination of diketones with aromatic amines a. [a] Reaction conditions: 1 (0.5 mmol, 1.0 equiv.), 2 (0.6 mmol, 1.1 equiv.), solvent (2.0 mL), catalyst (1.0 mol%), HCO2H (30.0 equiv.), under air, at 80 °C, for 12 h. [b] Isolated yield. [c] Determined by 1H NMR analysis.
Scheme 4. Ir-catalyzed reductive amination of diketones with aromatic amines a. [a] Reaction conditions: 1 (0.5 mmol, 1.0 equiv.), 2 (0.6 mmol, 1.1 equiv.), solvent (2.0 mL), catalyst (1.0 mol%), HCO2H (30.0 equiv.), under air, at 80 °C, for 12 h. [b] Isolated yield. [c] Determined by 1H NMR analysis.
Molecules 29 02565 sch004
Scheme 5. Scale-up experiment.
Scheme 5. Scale-up experiment.
Molecules 29 02565 sch005
Scheme 6. Mechanistic studies on Ir-catalyzed reductive amination. (a) The transformation of 3a1′ to 3a1. (b) Ir-catalyzed reductive amination of pentane-2,4-dione with aniline.
Scheme 6. Mechanistic studies on Ir-catalyzed reductive amination. (a) The transformation of 3a1′ to 3a1. (b) Ir-catalyzed reductive amination of pentane-2,4-dione with aniline.
Molecules 29 02565 sch006
Scheme 7. Proposed mechanism. The asterisk “*” represents the excited state.
Scheme 7. Proposed mechanism. The asterisk “*” represents the excited state.
Molecules 29 02565 sch007
Table 1. Condition optimization of the iridium-catalyzed successive reductive amination for the synthesis of N-phenyl-substituted pyrrolidine (3a1) a.
Table 1. Condition optimization of the iridium-catalyzed successive reductive amination for the synthesis of N-phenyl-substituted pyrrolidine (3a1) a.
Molecules 29 02565 i001
EntryCatalystSolventHCO2H (equiv.)Yield of 3a1 (%) bYield of 3a1′ (%) bdr  c
1TC-1toluene20415560:40
2TC-2toluene20595159:41
3TC-3toluene20425263:37
4TC-4toluene20484864:36
5TC-5toluene20445450:50
6TC-6toluene20484858:42
7TC-2DMF20<565--
8TC-2dioxane20n.d.38--
9TC-2THF20n.d.76--
10TC-2MeOH2087850:50
11TC-2acetone20n.d.23--
12TC-2H2O2080<571:29
13TC-2DMSO20<531--
14TC-2H2O1071<566:34
15TC-2H2O1584<571:29
16TC-2H2O2595<571:29
17TC-2H2O3096 (92) d<571:29
18 eTC-2H2O3096<571:29
19 fTC-2H2O3094<571:29
20 gTC-2H2O3060<550:50
21--H2O30n.d.n.d.--
22TC-2H2O--n.d.n.d.--
[a] Reaction conditions: 1a (0.5 mmol, 1.0 equiv.), 2a (0.6 mmol, 1.1 equiv.), solvent (2.0 mL), catalyst (1.0 mol%), HCO2H, under air, at 80 °C, for 12 h. [b] Determined by 1H NMR analysis using dimethyl terephthalate as the internal standard. The yield of 3a1 was determined with a mixture of cis-trans. [c] Determined by 1H NMR analysis. [d] Isolated yield. [e] At 100 °C. [f] 6 h. [g] At room temperature.
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

Liao, J.; Tong, J.; Liu, L.; Ouyang, L.; Luo, R. Construction of N-Aryl-Substituted Pyrrolidines by Successive Reductive Amination of Diketones via Transfer Hydrogenation. Molecules 2024, 29, 2565. https://doi.org/10.3390/molecules29112565

AMA Style

Liao J, Tong J, Liu L, Ouyang L, Luo R. Construction of N-Aryl-Substituted Pyrrolidines by Successive Reductive Amination of Diketones via Transfer Hydrogenation. Molecules. 2024; 29(11):2565. https://doi.org/10.3390/molecules29112565

Chicago/Turabian Style

Liao, Jianhua, Jinghui Tong, Liang Liu, Lu Ouyang, and Renshi Luo. 2024. "Construction of N-Aryl-Substituted Pyrrolidines by Successive Reductive Amination of Diketones via Transfer Hydrogenation" Molecules 29, no. 11: 2565. https://doi.org/10.3390/molecules29112565

Article Metrics

Back to TopTop