Next Article in Journal
Effect of Erosive Agents on Surface Characteristics of Nano-Fluorapatite Ceramic: An In-Vitro Study
Previous Article in Journal
Bridging the Chemical Profiles and Biological Effects of Spathodea campanulata Extracts: A New Contribution on the Road from Natural Treasure to Pharmacy Shelves
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

One-Pot Route from Halogenated Amides to Piperidines and Pyrrolidines

1
Department of Chemistry, Xihua University, Chengdu 610039, China
2
Asymmetric Synthesis and Chiral Technology Key Laboratory of Sichuan Province, Yibin 644000, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(15), 4698; https://doi.org/10.3390/molecules27154698
Submission received: 28 June 2022 / Revised: 17 July 2022 / Accepted: 19 July 2022 / Published: 22 July 2022
(This article belongs to the Section Organic Chemistry)

Abstract

:
Piperidine and pyrrolidine derivatives are important nitrogen heterocyclic structures with a wide range of biological activities. However, reported methods for their construction often face problems of requiring the use of expensive metal catalysts, highly toxic reaction reagents or hazardous reaction conditions. Herein, an efficient route from halogenated amides to piperidines and pyrrolidines was disclosed. In this method, amide activation, reduction of nitrile ions, and intramolecular nucleophilic substitution were integrated in a one-pot reaction. The reaction conditions were mild and no metal catalysts were used. The synthesis of a variety of N-substituted and some C-substituted piperidines and pyrrolidines became convenient, and good yields were obtained.

1. Introduction

Piperidine and pyrrolidine derivatives are important nitrogen heterocyclic structures with a wide range of biological activities [1,2,3,4,5]. For example, the first synthetic analgesic drug, pethidine, is a powerful analgesic still commonly used in clinical practice [6] (Figure 1, 1). Domperidone is a dopamine receptor antagonist that is used to treat digestive disorders [7] (Figure 1, 2). Levobupivacaine is clinically used as a long-acting local anesthetic that inhibits the release of γ-aminobutyric acid in the brain [8] (Figure 1, 3). Bepridil is a long-acting calcium antagonist clinically used to treat angina pectoris, arrhythmia, and hypertension [9] (Figure 1, 4). Buflomedil is a vasoactive drug with many pharmacological effects. It is widely used in the treatment of cerebrovascular and peripheral vascular diseases [10] (Figure 1, 5).
The reported methods for the construction of piperidines and pyrrolidines are mainly divided into two categories. The first type is the reduction method, in which expensive metal catalysts, such as palladium and platinum, are used to catalyze the hydrogenation reduction of pyridines or pyrroles [11,12,13,14,15]. Moreover, the reaction usually needs to be conducted at a high pressure of tens to hundreds of atmospheres. The other type is the cyclization reaction, in which cyclic amines are constructed by nucleophilic substitution of primary amines with dihaloalkanes or diols [16,17]. However, dihaloalkanes are strong alkylating agents with high genotoxicity, and the methods using diols also face the problem of requiring a high reaction temperature of over 200 °C.
Amides are a class of relatively stable carbonyl compounds that do not readily undergo reactions in contrast to acyl halides, anhydrides, and esters. However, a significant body of research on the selective activation of amides to achieve powerful transformations under mild conditions has emerged in recent decades [18,19,20,21,22,23,24]. In 2017, Huang’s group reported an interesting method that involves the amide activation-induced dehydracoupling of halogenated secondary amides with alkenes and NaBH4-triggered tandem cyclization reaction; the method is used to efficiently construct 2-allyl piperidines and pyrrolidines [25]. In 2021, our group disclosed a route from aryl ethylamide to fused indolizidines and quinolizidines [26]. In this method, the following are integrated in a one-pot reaction: amide activation, Bischler–Napieralski reaction (B–N reaction) [27,28], imine reduction, and intramolecular nucleophilic substitution. We found that if the reaction is always controlled at a lower temperature, then the B–N reaction will be inhibited. Through this, the main product of piperidines or pyrrolidines was obtained rather than the fused indolizidines and quinolizidines. In the present study, a one-pot route that involved the use of halogenated amides to construct piperidine and tetrahydropyrrole derivatives was disclosed (Scheme 1c).

2. Results and Discussion

Reaction conditions were optimized by using 5-chloro-N-(4-chlorolphenethyl)pentanamide (6a) as the starting material. Considering that pyridine class Lewis base plays an important role in the activation of amide, different kinds of base were first screened (Table 1, Entries 1–5). 2-Fluoropyridine (2-F-Py) was the most efficient (Table 1, Entry 4). When Lewis base was absent, the reaction still occurred, but the yield was reduced to a great extent (Table 1, Entry 6). Nevertheless, increasing the amount of 2-F-Py did not contribute to the improvement of yield (Table 1, Entry 7). The reaction temperature was subsequently investigated. When the reaction temperature was increased, the yield of the product decreased. (Table 1, Entries 8–9). This was due to the involving of B–N reaction at relatively higher temperatures, resulting in the production of some polycyclic byproducts. Different reductants, including KBH4, NaBH3CN, and NaBH(OAc)3, were then tested, but the outcomes were inferior to when NaBH4 was used (Table 1, Entries 10–12).
With the optimized reaction conditions defined, the scope of the reaction was investigated. N-phenethyl chloropentamides were tested first (Scheme 2, 7a7e). Substrates with electron-donating or electron-withdrawing substitutions in the aromatic nucleus all reacted smoothly to generate corresponding N-phenethyl piperidines. The reaction system was then applied to N-benzyl amides (Scheme 2, 7f7h). Various N-benzyl piperidines were obtained at moderate-to-good yields. Aliphatic chain and cyclic amide substrates adapted well under the established conditions and transformed into piperidine derivatives at good yields (Scheme 2, 7i7j). Halogenated butyramides were subsequently tested. Similar to piperidine derivatives, tetrahydropyrrole derivatives with versatile substitutions were also successfully synthesized (Scheme 2, 7k7t). We also attempted to synthesize C-substituted pyrrolidine. A 3-bromo-substituted pyrrolidine was successfully obtained in this method (Scheme 2, 7u).
We further attempted the synthesis of three-, four-, and seven-membered ring compounds. Unfortunately, none of these compounds were obtained, even if trace amounts were detected through mass spectrometry. The main products were uncycled secondary amines, which were confirmed by HRMS. Possibly, the ring tension made the reaction more difficult (Scheme 2, 7v). The synthesis of N-arylpyrrolidines was also difficult. The weaker nucleophilicity of arylamines than aliphatic amines may be the main factor that led to this result (Scheme 2, 7w). (See Supplementary Materials)
A plausible mechanism is illustrated in Scheme 3, with compound 7a as an example. Amide substrate (Scheme 3, 6a) was firstly activated by Tf2O to obtain a nitrilium ion [29,30] (Scheme 3, 8), which was reduced by sodium borohydride to obtain the imide ion (Scheme 3, 9). The imide ion 9 was further reduced by sodium borohydride to obtain the halogenated secondary amine [31], and this was followed by intramolecular nucleophilic substitution to obtain piperidine product (Scheme 3, 7a). In our previous work [26], the nitrilium ion 8 was readily attacked by electrons on the benzene ring and underwent the B–N reaction to form the imine ion (Scheme 3, 11) under 40 °C. Imine ion 11 was then subjected to reduction of the C=N bond (Scheme 3, 12) followed by intramolecular nucleophilic substitution to obtain a fused indolizidine product (Scheme 3, 13). However, in this work, it was difficult for the electrons on the benzene ring to attack imine ion 11 at a low temperature, so the B–N reaction was inhibited. As a result, a piperidine was obtained other than a fused indolizidine.

3. Materials and Methods

3.1. General Informations

All solvents were distilled from appropriate drying agents prior to use. CH2Cl2 was distilled over calcium hydride under a nitrogen atmosphere and stored over 4Å MS. Tf2O was distilled over phosphorous pentoxide (P2O5) and was stored for no more than a week before redistilling. Flash column chromatography was performed using silica gel (300–400 mesh). 1H NMR and 13C NMR (400 and 101 MHz, respectively) spectra were recorded on a Bruker 400 MHz NMR spectrometer in CDCl3. 1H NMR chemical shifts were reported in ppm (δ) relative to tetramethylsilane (TMS) with the solvent resonance employed as the internal standard (CDCl3, 7.26 ppm). 13C NMR chemical shifts were reported in ppm from TMS with the solvent resonance as the internal standard (CDCl3, 77.16 ppm). HRMS data were recorded on a SCIEX X500R QTOF HRMS apparatus.

3.2. General Procedure A for the Synthesis of Piperidines and Pyrrolidines

Into a dry 25 mL round-bottom flask equipped with a magnetic stirring bar, the following were added successively: a secondary amide (compound 6, 0.5 mmol, 1.0 equiv.), 10 mL of anhydrous CH2Cl2 and 2-F-Py (0.6 mmol, 1.2 equiv.) under an argon atmosphere. After being cooled to −78 °C, Tf2O (0.55 mmol, 1.1 equiv.) was added dropwise via a syringe, and the reaction was stirred for 30 min. Then, NaBH4 (1.0 mmol, 2 equiv.) and CH3OH (5 mL) were added under r.t. and stirred for additional 2 h. The reaction was quenched with a saturated aqueous solution of NaHCO3 (10 mL), and the mixture was extracted with dichloromethane (3 × 8 mL). The combined organic phase was dried over anhydrous Na2SO4, filtered and concentrated under reduced pressure. The residue was purified by flash chromatography on silica gel to give the corresponding compound.

4. Conclusions

In summary, we proposed a facile tandem protocol to construct piperidines and pyrrolidines. This method integrated amide activation, reduction of nitrile ions, and intramolecular nucleophilic substitution in a one-pot reaction. This method had mild reaction conditions and produced a variety of N-substituted and some C-substituted piperidines and pyrrolidines at good yields.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27154698/s1, 1H NMR data and spectrum, HMRS data, character for compound 7a7u; 13C NMR data and spectrum for compound 7u; structures of the amide substrates. References [32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48] are cited in the Supplementary Materials.

Author Contributions

Synthetic experiments, Q.S., S.W. and Y.L.; formal analysis, X.L. and X.W.; supervision, Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Science and Technology Department of Sichuan Province, grant number 2021YFG0290 and 2021YFG0291; Science and Technology Department of Yibin, grant number 2021YG02 and Xihua University, grant number Z202030.

Data Availability Statement

The data presented in this study are available in Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Some of the compounds described in this article are available from the authors.

References

  1. Dragutan, I.; Dragutan, V.; Demonceau, A. Targeted Drugs by Olefin Metathesis: Piperidine-Based Iminosugars. RSC Adv. 2012, 2, 719–736. [Google Scholar] [CrossRef]
  2. Goel, P.; Alam, O.; Naim, M.J.; Nawaz, F.; Iqbal, M.; Alam, M.I. Recent Advancement of Piperidine Moiety in Treatment of Cancer—A Review. Eur. J. Med. Chem. 2018, 157, 480–502. [Google Scholar] [CrossRef] [PubMed]
  3. Islam, M.T.; Mubarak, M.S. Pyrrolidine Alkaloids and Their Promises in Pharmacotherapy. Adv. Tradit. Med. 2020, 20, 13–22. [Google Scholar] [CrossRef]
  4. Rathore, A.; Asati, V.; Kashaw, S.K.; Agarwal, S.; Parwani, D.; Bhattacharya, S.; Mallick, C. The Recent Development of Piperazine and Piperidine Derivatives as Antipsychotic Agents. MINI-Rev. Med. Chem. 2021, 21, 362–379. [Google Scholar] [CrossRef]
  5. Li Petri, G.; Raimondi, M.V.; Spano, V.; Holl, R.; Barraja, P.; Montalbano, A. Pyrrolidine in Drug Discovery: A Versatile Scaffold for Novel Biologically Active Compounds. Top. Curr. Chem. 2021, 379, 34. [Google Scholar] [CrossRef]
  6. Lascelles, X.B.D.; Cripps, J.P.; Jones, A.; Waterman, E.A. Post-Operative Central Hypersensitivity and Pain: The Pre-Emptive Value of Pethidine for Ovariohysterectomy. Pain 1997, 73, 461–471. [Google Scholar] [CrossRef]
  7. Zhang, C.-H.; Zhao, B.-X.; Huang, Y.; Wang, Y.; Ke, X.-Y.; Zhao, B.-J.; Zhang, X.; Zhang, Q. A Novel Domperidone Hydrogel: Preparation, Characterization, Pharmacokinetic, and Pharmacodynamic Properties. J. Drug Deliv. 2011, 2011, 841054. [Google Scholar] [CrossRef]
  8. Robinson, A.P.; Lyons, G.R.; Wilson, R.C.; Gorton, H.J.; Columb, M.O. Levobupivacaine for Epidural Analgesia in Labor: The Sparing Effect of Epidural Fentanyl. Anesth. Analg. 2001, 92, 410–414. [Google Scholar] [CrossRef]
  9. Takahara, A.; Nakamura, Y.; Sugiyama, A. Beat-to-Beat Variability of Repolarization Differentiates the Extent of Torsadogenic Potential of Multi Ion Channel-Blockers Bepridil and Amiodarone. Eur. J. Pharmacol. 2008, 596, 127–131. [Google Scholar] [CrossRef]
  10. Bourguignon, L.; Ducher, M.; Matanza, D.; Bleyzac, N.; Uhart, M.; Odouard, E.; Maire, P.; Goutelle, S. The Value of Population Pharmacokinetics and Simulation for Postmarketing Safety Evaluation of Dosing Guidelines for Drugs with a Narrow Therapeutic Index: Buflomedil as a Case Study: Prediction of Buflomedil Overdosing in Elderly Patients. Fundam. Clin. Pharmacol. 2012, 26, 279–285. [Google Scholar] [CrossRef]
  11. Adkins, H.; Kuick, L.F.; Farlow, M.; Wojcik, B. Hydrogenation of Derivatives of Pyridine. J. Am. Chem. Soc. 1934, 56, 2425–2428. [Google Scholar] [CrossRef]
  12. Wang, Y.; Cui, X.; Deng, Y.; Shi, F. Catalytic Hydrogenation of Aromatic Rings Catalyzed by Pd/NiO. RSC Adv. 2014, 4, 2729–2732. [Google Scholar] [CrossRef]
  13. Piras, L.; Genesio, E.; Ghiron, C.; Taddei, M. Microwave-Assisted Hydrogenation of Pyridines. Synlett 2008, 2008, 1125–1128. [Google Scholar]
  14. Karakulina, A.; Gopakumar, A.; Akçok, İ.; Roulier, B.L.; LaGrange, T.; Katsyuba, S.A.; Das, S.; Dyson, P.J. A Rhodium Nanoparticle–Lewis Acidic Ionic Liquid Catalyst for the Chemoselective Reduction of Heteroarenes. Angew. Chem. 2016, 128, 300–304. [Google Scholar] [CrossRef]
  15. Hegedűs, L.; Szőke-Molnár, K.; Sajó, I.E.; Srankó, D.F.; Schay, Z. Poisoning and Reuse of Supported Precious Metal Catalysts in the Hydrogenation of N-Heterocycles Part I: Ruthenium-Catalysed Hydrogenation of 1-Methylpyrrole. Catal. Lett. 2018, 148, 1939–1950. [Google Scholar] [CrossRef]
  16. Fujita, K.; Enoki, Y.; Yamaguchi, R.; Moura-Letts, G.; Curran, D.P. Iridium-Catalyzed N-Heterocyclization of Primary Amines with Diols: N-Benzylpiperidine. Org. Synth. 2006, 83, 217–221. [Google Scholar]
  17. Berger, M.L.; Schweifer, A.; Rebernik, P.; Hammerschmidt, F. NMDA Receptor Affinities of 1,2-Diphenylethylamine and 1-(1,2-Diphenylethyl)Piperidine Enantiomers and of Related Compounds. Bioorg. Med. Chem. 2009, 17, 3456–3462. [Google Scholar] [CrossRef]
  18. Falmagne, J.-B.; Escudero, J.; Taleb-Sahraoui, S.; Ghosez, L. Cyclobutanone and Cyclobutenone Derivatives by Reaction of Tertiary Amides with Alkenes or Alkynes. Angew. Chem. Int. Ed. Engl. 1981, 20, 879–880. [Google Scholar] [CrossRef]
  19. Movassaghi, M.; Hill, M.D.; Ahmad, O.K. Direct Synthesis of Pyridine Derivatives. J. Am. Chem. Soc. 2007, 129, 10096–10097. [Google Scholar] [CrossRef]
  20. Movassaghi, M.; Hill, M.D. Synthesis of Substituted Pyridine Derivatives via the Ruthenium-Catalyzed Cycloisomerization of 3-Azadienynes. J. Am. Chem. Soc. 2006, 128, 4592–4593. [Google Scholar] [CrossRef]
  21. Huang, P.-Q. Direct Transformations of Amides: Tactics and Recent Progress. Acta Chim. Sin. 2018, 76, 357. [Google Scholar] [CrossRef]
  22. Adler, P.; Teskey, C.J.; Kaiser, D.; Holy, M.; Sitte, H.H.; Maulide, N. α-Fluorination of Carbonyls with Nucleophilic Fluorine. Nat. Chem. 2019, 11, 329–334. [Google Scholar] [CrossRef]
  23. Huang, X.-Z.; Gao, L.-H.; Huang, P.-Q. Enantioselective Total Syntheses of (+)-Stemofoline and Three Congeners Based on a Biogenetic Hypothesis. Nat. Commun. 2020, 11, 5314. [Google Scholar] [CrossRef]
  24. Kaiser, D.; Bauer, A.; Lemmerer, M.; Maulide, N. Amide Activation: An Emerging Tool for Chemoselective Synthesis. Chem. Soc. Rev. 2018, 47, 7899–7925. [Google Scholar] [CrossRef] [Green Version]
  25. Huang, P.-Q.; Huang, Y.-H.; Wang, S.-R. One-Pot Synthesis of N-Heterocycles and Enimino Carbocycles by Tandem Dehydrative Coupling–Reductive Cyclization of Halo-Sec-Amides and Dehydrative Cyclization of Olefinic Sec-Amides. Org. Chem. Front. 2017, 4, 431–444. [Google Scholar] [CrossRef] [Green Version]
  26. Song, Q.; Liu, Y.; Cai, L.; Cao, X.; Qian, S.; Wang, Z. One-Pot Tandem Route to Fused Indolizidines and Quinolizidines: Application in the Synthesis of Alkaloids and Bioactive Compounds. Chin. Chem. Lett. 2021, 32, 1713–1716. [Google Scholar] [CrossRef]
  27. Bischler, A.; Napieralski, B. Zur Kenntniss Einer Neuen Isochinolinsynthese. Ber. Dtsch. Chem. Ges. 1893, 26, 1903–1908. [Google Scholar] [CrossRef] [Green Version]
  28. Bergstrom, F. Heterocyclic Nitrogen Compounds. Part IIA. Hexacyclic Compounds: Pyridine, Quinoline, and Isoquinoline. Chem. Rev. 1944, 35, 77–277. [Google Scholar] [CrossRef]
  29. Charette, A.B.; Grenon, M. Spectroscopic Studies of the Electrophilic Activation of Amides with Triflic Anhydride and Pyridine. Can. J. Chem. 2001, 79, 1694–1703. [Google Scholar] [CrossRef]
  30. Ye, J.-L.; Zhu, Y.-N.; Geng, H.; Huang, P.-Q. Metal-Free Synthesis of Quinolines by Direct Condensation of Amides with Alkynes: Revelation of N-Aryl Nitrilium Intermediates by 2D NMR Techniques. Sci. China Chem. 2018, 61, 687–694. [Google Scholar] [CrossRef]
  31. Xiang, S.-H.; Xu, J.; Yuan, H.-Q.; Huang, P.-Q. Amide Activation by Tf2O: Reduction of Amides to Amines by NaBH4 under Mild Conditions. Synlett 2010, 2010, 1829–1832. [Google Scholar]
  32. Zhang, X.-Y.; Zhai, D.-D.; Liu, Y.-F.; Guan, B.-T. A Potassium Magnesiate Complex: Synthesis, Structure and Catalytic Intermolecular Hydroamination of Styrenes. J. Organomet. Chem. 2022, 961, 122254. [Google Scholar] [CrossRef]
  33. Mulks, F.F.; Bole, L.J.; Davin, L.; Hernán-Gómez, A.; Kennedy, A.; García-Álvarez, J.; Hevia, E. Ambient Moisture Accelerates Hydroamination Reactions of Vinylarenes with Alkali-Metal Amides under Air. Angew. Chem. Int. Ed. 2020, 59, 19021–19026. [Google Scholar] [CrossRef] [PubMed]
  34. Otsuka, M.; Yokoyama, H.; Endo, K.; Shibata, T. Ru-Catalyzed β-Selective and Enantioselective Addition of Amines to Styrenes Initiated by Direct Arene-Exchange. Org. Biomol. Chem. 2012, 10, 3815–3818. [Google Scholar] [CrossRef]
  35. Pandey, P.; Bera, J.K. Hydrosilylative Reduction of Primary Amides to Primary Amines Catalyzed by a Terminal [Ni–OH] Complex. Chem. Commun. 2021, 57, 9204–9207. [Google Scholar] [CrossRef]
  36. Hammerstad, T.A.; Hegde, P.V.; Wang, K.J.; Aldrich, C.C. Chemoselective Reduction of Tertiary Amides by 1, 3-Diphenyldisiloxane (DPDS). Synthesis 2022, 54, 2205–2212. [Google Scholar] [CrossRef]
  37. Barger, C.J.; Dicken, R.D.; Weidner, V.L.; Motta, A.; Lohr, T.L.; Marks, T.J. La [N (SiMe3) 2] 3-Catalyzed Deoxygenative Reduction of Amides with Pinacolborane. Scope and Mechanism. J. Am. Chem. Soc. 2020, 142, 8019–8028. [Google Scholar] [CrossRef]
  38. Talybov, A.; Mamedbeili, E.; Abbasov, V.; Kochetkov, K. Synthesis and Properties of Pentane Amino Derivatives. Russ. J. Gen. Chem. 2010, 80, 2455–2459. [Google Scholar] [CrossRef]
  39. Zeng, H.; Cao, D.; Qiu, Z.; Li, C.-J. Palladium-Catalyzed Formal Cross-Coupling of Diaryl Ethers with Amines: Slicing the 4-O-5 Linkage in Lignin Models. Angew. Chem. 2018, 130, 3814–3819. [Google Scholar] [CrossRef]
  40. Germain, S.; Schulz, E.; Hannedouche, J. Anti-Markovnikov Hydroamination of Aromatic Alkenes with Secondary Amines Catalyzed by Easily Accessible Yttrium Complexes. ChemCatChem 2014, 6, 2065–2073. [Google Scholar] [CrossRef]
  41. Taylor, N.J.; Emer, E.; Preshlock, S.; Schedler, M.; Tredwell, M.; Verhoog, S.; Mercier, J.; Genicot, C.; Gouverneur, V. Derisking the Cu-Mediated 18F-Fluorination of Heterocyclic Positron Emission Tomography Radioligands. J. Am. Chem. Soc. 2017, 139, 8267–8276. [Google Scholar] [CrossRef]
  42. Relitti, N.; Federico, S.; Pozzetti, L.; Butini, S.; Lamponi, S.; Taramelli, D.; D’Alessandro, S.; Martin, R.E.; Shafik, S.H.; Summers, R.L.; et al. Synthesis and Biological Evaluation of Benzhydryl-Based Antiplasmodial Agents Possessing Plasmodium Falciparum Chloroquine Resistance Transporter (PfCRT) Inhibitory Activity. Eur. J. Med. Chem. 2021, 215, 113227. [Google Scholar] [CrossRef]
  43. Nersesian, D.L.; Black, L.A.; Miller, T.R.; Vortherms, T.A.; Esbenshade, T.A.; Hancock, A.A.; Cowart, M.D. In Vitro SAR of Pyrrolidine-Containing Histamine H3 Receptor Antagonists: Trends across Multiple Chemical Series. Bioorg. Med. Chem. Lett. 2008, 18, 355–359. [Google Scholar] [CrossRef]
  44. Valverde, E.A.; Romero, A.H.; Acosta, M.E.; Gamboa, N.; Henriques, G.; Rodrigues, J.R.; Ciangherotti, C.; López, S.E. Synthesis, β-Hematin Inhibition Studies and Antimalarial Evaluation of New Dehydroxy Isoquine Derivatives against Plasmodium Berghei: A Promising Antimalarial Agent. Eur. J. Med. Chem. 2018, 148, 498–506. [Google Scholar] [CrossRef]
  45. Amberchan, G.; Snelling, R.A.; Moya, E.; Landi, M.; Lutz, K.; Gatihi, R.; Singaram, B. Reaction of Diisobutylaluminum Borohydride, a Binary Hydride, with Selected Organic Compounds Containing Representative Functional Groups. J. Org. Chem. 2021, 86, 6207–6227. [Google Scholar] [CrossRef]
  46. Varjosaari, S.E.; Skrypai, V.; Suating, P.; Hurley, J.J.; Lio, A.M.D.; Gilbert, T.M.; Adler, M.J. Simple Metal-Free Direct Reductive Amination Using Hydrosilatrane to Form Secondary and Tertiary Amines. Adv. Synth. Catal. 2017, 359, 1872–1878. [Google Scholar] [CrossRef]
  47. Kadyrov, R.; Moebus, K. Reductive Alkylation of Amines with Carboxylic Ortho Esters. Adv. Synth. Catal. 2020, 362, 3352–3357. [Google Scholar] [CrossRef]
  48. Elmer, L.-M.; Kehr, G.; Daniliuc, C.G.; Siedow, M.; Eckert, H.; Tesch, M.; Studer, A.; Williams, K.; Warren, T.H.; Erker, G. The Chemistry of a Non-Interacting Vicinal Frustrated Phosphane/Borane Lewis Pair. Chem. Eur. J. 2017, 23, 6056–6068. [Google Scholar] [CrossRef]
Figure 1. Drugs containing piperidine or pyrrolidine structure.
Figure 1. Drugs containing piperidine or pyrrolidine structure.
Molecules 27 04698 g001
Scheme 1. Comparison of previous works and this work. (a) Huang’s work; (b) Our previous work; (c) This work.
Scheme 1. Comparison of previous works and this work. (a) Huang’s work; (b) Our previous work; (c) This work.
Molecules 27 04698 sch001
Scheme 2. Scope of this method a: For each compound, the isolated yield is given as a percentage. a Uniform reaction conditions unless otherwise noted: Compound 6 (0.5 mmol), Tf2O (1.1 equiv.), 2-F-Py (1.2 equiv.), CH2Cl2 (5 mL), Ar, −78 °C, 30 min → MeOH (5 mL), NaBH4 (2.0 equiv.), r.t., 2 h.
Scheme 2. Scope of this method a: For each compound, the isolated yield is given as a percentage. a Uniform reaction conditions unless otherwise noted: Compound 6 (0.5 mmol), Tf2O (1.1 equiv.), 2-F-Py (1.2 equiv.), CH2Cl2 (5 mL), Ar, −78 °C, 30 min → MeOH (5 mL), NaBH4 (2.0 equiv.), r.t., 2 h.
Molecules 27 04698 sch002
Scheme 3. The proposed mechanism.
Scheme 3. The proposed mechanism.
Molecules 27 04698 sch003
Table 1. Optimization of reaction conditions a.
Table 1. Optimization of reaction conditions a.
Molecules 27 04698 i001
EntryBaseReaction T (°C)ReductantYield (%) b
1Pyridine (1.2 equiv.)-−78NaBH428
23,5-dimethylpyridine (1.2 equiv.)−78NaBH432
32-Cl-Py (1.2 equiv.)−78NaBH450
42-F-Py (1.2 equiv.)−78NaBH490
52-I-Py (1.2 equiv.)−78NaBH440
6-−78NaBH412
72-F-Py (2.0 equiv.)−78NaBH485
82-F-Py (1.2 equiv.)−30NaBH473
92-F-Py (1.2 equiv.)0NaBH466
102-F-Py (1.2 equiv.)−78KBH475
112-F-Py (1.2 equiv.)−78NaBH3CN40
122-F-Py (1.2 equiv.)−78NaBH(OAc)3trace
a Reaction conditions: 6a (0.5 mmol), Tf2O (1.1 equiv.), base, CH2Cl2 (5 mL), Ar, reaction temp., 30 min → reductant (2.0 equiv.), MeOH (5 mL), r.t., 2 h. b Yields were determined by 1H NMR with 1,1,2,2-tetrachloroethane as the internal standard.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Song, Q.; Wang, S.; Lei, X.; Liu, Y.; Wen, X.; Wang, Z. One-Pot Route from Halogenated Amides to Piperidines and Pyrrolidines. Molecules 2022, 27, 4698. https://doi.org/10.3390/molecules27154698

AMA Style

Song Q, Wang S, Lei X, Liu Y, Wen X, Wang Z. One-Pot Route from Halogenated Amides to Piperidines and Pyrrolidines. Molecules. 2022; 27(15):4698. https://doi.org/10.3390/molecules27154698

Chicago/Turabian Style

Song, Qiao, Sheng Wang, Xiangui Lei, Yan Liu, Xin Wen, and Zhouyu Wang. 2022. "One-Pot Route from Halogenated Amides to Piperidines and Pyrrolidines" Molecules 27, no. 15: 4698. https://doi.org/10.3390/molecules27154698

Article Metrics

Back to TopTop