Next Article in Journal
Interactions of Tea-Derived Catechin Gallates with Bacterial Pathogens
Next Article in Special Issue
An Efficient Greener Approach for N-acylation of Amines in Water Using Benzotriazole Chemistry
Previous Article in Journal
In Vitro and in Vivo Activity of mTOR Kinase and PI3K Inhibitors Against Leishmania donovani and Trypanosoma brucei
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N-Acetylation of Amines in Continuous-Flow with Acetonitrile—No Need for Hazardous and Toxic Carboxylic Acid Derivatives

by
György Orsy
1,2,
Ferenc Fülöp
1,3,* and
István M. Mándity
2,4,*
1
Institute of Pharmaceutical Chemistry, University of Szeged, Eötvös u. 6, H-6720 Szeged, Hungary
2
MTA TTK Lendület Artificial Transporter Research Group, Institute of Materials and Environmental Chemistry, Research Center for Natural Sciences, Hungarian Academy of Sciences, Magyar Tudosok krt. 2, 1117 Budapest, Hungary
3
Research Group of Stereochemistry of the Hungarian Academy of Sciences, Dóm tér 8, H-6720 Szeged, Hungary
4
Department of Organic Chemistry, Faculty of Pharmacy, Semmelweis University, Hőgyes Endre u. 7, H-1092 Budapest, Hungary
*
Authors to whom correspondence should be addressed.
Molecules 2020, 25(8), 1985; https://doi.org/10.3390/molecules25081985
Submission received: 25 March 2020 / Revised: 17 April 2020 / Accepted: 21 April 2020 / Published: 23 April 2020

Abstract

:
A continuous-flow acetylation reaction was developed, applying cheap and safe reagent, acetonitrile as acetylation agent and alumina as catalyst. The method developed utilizes milder reagent than those used conventionally. The reaction was tested on various aromatic and aliphatic amines with good conversion. The catalyst showed excellent reusability and a scale-up was also carried out. Furthermore, a drug substance (paracetamol) was also synthesized with good conversion and yield.

Graphical Abstract

1. Introduction

N-acetylation is a widely used chemical reaction in general organic chemistry to build an acetyl functional group on an amine compound [1,2,3,4]. The use of the acetyl functional group is widespread, including drug research, the preparation of pharmaceuticals, polymer chemistry and agrochemical applications [5,6,7,8,9,10]. It can be utilized as a protecting group in many organic reactions and also in peptide synthesis [11]. In addition, it plays a major regulatory role in post-translational protein modification and regulation of DNA expression in all life forms [12,13].
In general, common acetylation reagents such as acetic anhydride and acetyl chloride, are easily accessible in chemical laboratories. Even the most sustainable technologies utilize these reagents in combination with various Lewis acids [14,15,16] and/or in neat form [17]. Nevertheless, the utilization of acetic anhydride and acetyl chloride have various drawbacks. Both reagents are major irritants and acetyl chloride is considered to be a genotoxic agent [18]. As such, the elimination of their use is of considerable current interest.
Flow chemistry technology is widely used in many synthetic organic reactions at both laboratory and industrial scale [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. There are a number of benefits of using continuous-flow (CF) chemistry. A wide range of reactions are much faster in flow processes, and fewer substrates and reagents are required [36,37,38,39,40]. Furthermore, more efficient and selective reaction can be carried out in continuous systems than in regular batch operations [41,42,43,44,45,46]. Additionally, flow reaction conditions can enable reaction routes that would otherwise only be feasible under regular batch conditions, e.g., higher temperature and pressure than can be used under safe conditions [47,48,49,50,51,52,53,54,55,56,57].
Whilst acetonitrile is a common solvent and is generally used in various fields of chemistry as eluent [41] and polar aprotic organic solvent [58], it is rarely used as reagent in organic chemistry. A few studies on acetonitrile as an acylation agent have been reported [59,60,61,62,63,64] thus far, for example Saikia et al., [65] presenting an unusual attempt to synthesize N-acylated aromatic amines. Acetonitrile was utilized as reagent and solvent with several Lewis acids (e.g., Cu(OAc)2, Mn(OAc)2, FeCl3, InCl3). The most promising catalyst was the trimethylsilyl iodide (TMSI), which activated the acetonitrile by co-ordination. On the other hand, Brahmayya et al. [66] developed a new method for preparing N-acetamides with metal-free sulfonated reduced graphene oxide catalyst under sonication.
Herein, we present the efficient utilization of acetonitrile for acetylation. We describe a selective and environmentally friendly CF acetylation of aromatic and aliphatic amines by the use of a low-cost and environmentally friendly Lewis acid catalyst as alumina in acetonitrile solvent with excellent conversions. These observations may open a new and green procedure for the synthesis of acetylated aromatic and aliphatic amines.

2. Results and Discussion

The reactions were carried out in a home-made CF reactor (See Supplementary Material Figure S23). Our equipment consists of an HPLC pump that transports the substrate dissolved in acetonitrile. The solution is feed into a fillable HPLC column where the reaction occurs. The column is filled with solid catalyst. Additionally, there is a GC oven and an in-line back pressure regulator in the system to control the temperature and pressure of the reaction. A schematic outline of the reactor used in this study is shown in Figure 1.
In order to find the most useful solid Lewis acid catalyst for the flow synthesis, a reagent screen was carried out. Aniline, as a test compound was utilized, and acetonitrile was used both as solvent and acyl donor. The complete study of Lewis acids is shown in Table 1. The study shows that the most promising Lewis acid was the aluminum(III) oxide, other catalysts offered lower yields or no amide product formation was observed. Thus, the further reaction parameter optimization was carried out with aluminum(III) oxide as solid catalyst.
The starting material was dissolved in acetonitrile in a concentration of 100 mM, using aluminum(III) oxide powder as catalyst at a temperature of 25 °C and a pressure of 10–100 bar with 0.1 mL min−1 flow rate and 27 min residence time. The reaction did not show any pressure dependency and a moderate conversion of 27% was observed without any real pressure dependence. The effect of temperature on the conversion rate was tested at 50 bar pressure. Raising the temperature to 100 °C initiated the product formation with 53% conversion. By applying a higher temperature, the conversion of the reaction improved remarkably. It was found that the optimal temperature is 200 °C. Further increase of the temperature resulted in lower conversion values. In this case, the acetonitrile was in its supercritical state (Tc = 275 °C, Pc = 48 bar) due to the significant solvent expansion produced under supercritical conditions. This is a likely explanation of the decrease of the observed yield above 275 °C. The results obtained for the temperature dependence are summarized in (Figure 2a). The effect of pressure on reaction outcome was also tested (Figure 2b). The results indicate that a modest pressure (50 bar) is needed to increase the reaction conversion, while further pressure increase did not influence the reaction outcome significantly. The flow rate was tested too (Figure 2c). The results show that the optimum flow rate is 0.1 mL min−1. Any increase in the flow rate above this value resulted in decreased conversion. The effect of concentration on the reaction outcome was also tested (Figure 2d). The results show that increased concentration results in the decreased conversion.
Under the optimized conditions, the scope of the reaction was explored by a variety of aniline derivatives (Table 2). The anilines containing either electron-donating or electron-withdrawing groups were selected. It must be noted that all reactions were carried out in a single run and the products were analyzed by 1H and 13C NMR spectroscopy. Column chromatography purification of the product was only needed for compound 6. For the others, only a simple evaporation of acetonitrile was required.
Importantly, the hydroxyl group possessing aniline 3 was acetylated with excellent yield and the drug substance paracetamol 4 with quantitative yield after a simple recrystallization. Lower yield was observed for 4-methoxyaniline (5), and the acetylated product was isolated after column chromatography with a mere 51% yield. For the halogen atom possessing anilines (7, 9 and 11) excellent yields were achieved. For nitroanilines (13,15 and 17) no conversion was observed and only the starting material was isolated. This fact might be explained by the highly electron-withdrawing nature of the nitro substituent, which reduces the nucleophilicity of the amino group. Besides aromatic amines, aliphatic primary and secondary amines were also tested. The primary benzylic amine (23) was converted to the corresponding acetamide with excellent yield. When secondary amine piperidine (19) and morpholine (21) were tested, the acetylated derivatives (20, 22) were isolated with quantitative yields. The use of this reaction in stereoselective reaction was also tested with acetylation reaction carried out for the two enantiomers of 1-phenylethanamine (25 and 27). For both enantiomers the acetylated derivative was achieved with quantitative yield and the complete retention of the enatiomerical purity. This was investigated by optical rotation and was found to be identical to literature data.
The catalyst reusability was also tested. It was an important finding that the activity of the catalyst did not decrease significantly until 10 cycles and one cycle was carried out with 20 mg of benzylamine. This excellent result [67] opened the way to scale up the reaction, which was tested with the same reaction. The acetylation process could be scaled up to 2 g of benzylamine without significant decrease of the conversion. The product (N-benzylacetamide, 24) was isolated with 94% yield after a simple recrystallization. The reaction was completed within 28 hours. This is considerably faster than what has been reported with already known technologies. Results are shown in Figure 3.
These results can be explained by the proposed reaction mechanism (Scheme 1) which relies on literature data [63,64,65,66]. A key step is the coordination of the lone electron pair of the nitrogen atom of the cyanide group, which yields a positive charge. Due to mesomeric structures, the positive charge might be localized on the carbon atom of the cyanide group. This positively charged carbon atom might be attacked by the lone pair electron of the amine yielding amidine, which is further hydrolyzed to provide the acetamide as shown in Scheme 1. The origin of the water is the residual water content of the solvent and the alumina. The addition of extra amount of water decreased the conversion of the reaction.

3. Materials and Methods

3.1. General

All solvents and reagents were of analytical grade and used directly without further purification. Fe2O3, Boric acid, AlCl3, Al2O3 (for chromatography, activated, neutral, Brockmann I, 50–200 µm, 60 A) catalysts used in this study were purchased from Sigma-Aldrich (Budapest, Hungary), while Acetonitrile (100, 0%) was HPLC LC MS-grade solvents from VWR International (Debrecen, Hungary).

3.2. General Aspects of the CF Acetylation

The CF acetylation reactions were carried out in a home-made flow reactor consisting of an HPLC pump (Jasco PU-987 Intelligent Prep. Pump), a stainless steel HPLC column as catalyst bed (internal dimensions 250mm L × 4.6 ID × ¼ in OD), a stainless steel preheating coil (internal diameter 1 mm and length 30 cm) and a commercially available backpressure regulator (Thalesnano back pressure module 300™, Budapest, Hungary, to a maximum of 300 bar). Parts of the system were connected with stainless steel tubing (internal diameter 1 mm). The HPLC column was charged with 4 g of the alumina catalyst. It was then placed into a GC oven unit (Carlo Erba HR 5300 up to maximum a 350 °C). For the CF reactions, 100 mM solution of the appropriate starting material was prepared in acetonitrile. The solution was homogenized by sonication for 5 min and then pumped through the CF reactor under the set conditions. After the completion of the reaction, the reaction mixture was collected and the rest solvent was evaporated by a vacuum rotary evaporator.

3.3. Product Analysis

The products obtained were characterized by NMR spectroscopy. 1H-NMR and 13C-NMR spectra were recorded on a Bruker Avance DRX 400 spectrometer (Bruker AVANCE, Billerica, MA, USA), in DMSO-d6 as solvent, at 400.1 MHz. Chemical shifts (δ) are expressed in ppm and are internally referenced (1H NMR: 2.50 ppm in DMSO-d6). Conversion was determined via the 1H-NMR spectra of the crude materials. A PerkinElmer 341 polarimeter (PerkinElmer, Boston, MA, USA) was used for the determination of optical rotations.

4. Conclusions

We have developed a sustainable CF process for the selective N-acetylation of various amines. The obtained chemical process is time and cost efficient, as it utilizes cheap reagent and catalyst and considerably faster with a residence time of only 27 min. Importantly, the well-known, cheap and non-toxic Lewis acid alumina was used as catalyst. Moreover, the acetylation reagent was acetonitrile, which is an industrial side-product with a modest price and it is considerably milder than those for the regular carboxylic acid derivatives used for acetylation, e.g., acetyl chloride, acetic anhydride. In general, under the optimized conditions good or excellent conversions were observed for the aromatic amines, except for the nitro substituted compounds, where no conversion was reached. Importantly, the painkiller drug substance paracetamol was gained with high yield. Mainly complete conversions were also observed for primary and secondary aliphatic amines. The use of this acetylation in stereoselective reactions was also tested and during the course of the reaction no racemization was observed for either enantiomer of enantiomerically pure 1-phenylethanamine.

Supplementary Materials

The following are available online, general experimetal data, NMR spectra, image of the reactor used.

Author Contributions

Conceptualization, I.M.M. and F.F.; Investigation, G.O.; Supervision, I.M.M. and F.F.; Writing—Original draft, G.O.; Writing—Review & editing, I.M.M. and F.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Hungarian Ministry for Innovation and Technology, grant number 2018-1.2.1-NKP-2018-00005 and by National Research, Development and Innovation Fund of Hungary, grant number 2018-1.2.1-NKP.

Acknowledgments

The Lendület grant from the Hungarian Academy of Sciences is gratefully acknowledged. This work was completed in the ELTE Thematic Excellence Programme supported by the Hungarian Ministry for Innovation and Technology. Project no. 2018-1.2.1-NKP-2018-00005 has been implemented with the support provided from the National Research, Development and Innovation Fund of Hungary, financed under the 2018-1.2.1-NKP funding scheme.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Amić, A.; Molnar, M. An Improved and Efficient N-acetylation of Amines Using Choline Chloride Based Deep Eutectic Solvents. Org. Prep. Proc. Int. 2017, 49, 249–257. [Google Scholar] [CrossRef]
  2. Biswas, R.; Mukherjee, A. Introducing the Concept of Green Synthesis in the Undergraduate Laboratory: Two-Step Synthesis of 4-Bromoacetanilide from Aniline. J. Chem. Ed. 2017, 94, 1391–1394. [Google Scholar] [CrossRef]
  3. Tajbakhsh, M.; Hosseinzadeh, R.; Alinezhad, H.; Rezaee, P.; Tajbakhsh, M. TiO2 NPs as Catalyst for N-Formylation and N-Acetylation of Amines under Solvent-Free Conditions. Lett. Org. Chem. 2013, 10, 657–663. [Google Scholar] [CrossRef] [Green Version]
  4. Kim, D.H. Acetic anhydride as a synthetic reagent. J. Het. Chem. 1976, 13, 179–194. [Google Scholar] [CrossRef]
  5. Baumann, M.; Baxendale, I.R. The synthesis of active pharmaceutical ingredients (APIs) using continuous flow chemistry. Beilstein J. Org. Chem. 2015, 11, 1194–1219. [Google Scholar] [CrossRef] [Green Version]
  6. Britton, J.; Raston, C.L. Multi-step continuous-flow synthesis. Chem. Soc. Rev. 2017, 46, 1250–1271. [Google Scholar] [CrossRef] [Green Version]
  7. Escobar Carlos, A.; Orellana-Vera, J.; Vega, A.; Sicker, D.; Sieler, J. Regioselective N-Acetylation of 4-(2-Hydroxyphenyl)-2-phenyl-2,3-dihydro-1H-1,5-benzodiazepine Using Protection by an Intramolecular Hydrogen Bond. Z. Für Nat. B 2009, 64, 969–972. [Google Scholar] [CrossRef] [Green Version]
  8. Farina, V.; Reeves, J.T.; Senanayake, C.H.; Song, J.J. Asymmetric Synthesis of Active Pharmaceutical Ingredients. Chem. Rev. 2006, 106, 2734–2793. [Google Scholar] [CrossRef]
  9. Nakazono, K.; Fukasawa, K.; Sato, T.; Koyama, Y.; Takata, T. Synthesis of acetylene-functionalized [2]rotaxane monomers directed toward side chain-type polyrotaxanes. Polym. J. 2010, 42, 208–215. [Google Scholar] [CrossRef] [Green Version]
  10. Xu, G.; Tang, D.; Gai, Y.; Wang, G.; Kim, H.; Chen, Z.; Phan, L.T.; Or, Y.S.; Wang, Z. An Efficient Large-Scale Synthesis of EDP-420, a First-in-Class Bridged Bicyclic Macrolide (BBM) Antibiotic Drug Candidate. Org. Proc. Res. Dev. 2010, 14, 504–510. [Google Scholar] [CrossRef]
  11. Isidro-Llobet, A.; Álvarez, M.; Albericio, F. Amino Acid-Protecting Groups. Chem. Rev. 2009, 109, 2455–2504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Hansen, B.K.; Gupta, R.; Baldus, L.; Lyon, D.; Narita, T.; Lammers, M.; Choudhary, C.; Weinert, B.T. Analysis of human acetylation stoichiometry defines mechanistic constraints on protein regulation. Nat. Commun. 2019, 10, 1055. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, Y.; Zhou, F.; Bai, M.; Liu, Y.; Zhang, L.; Zhu, Q.; Bi, Y.; Ning, G.; Zhou, L.; Wang, X. The pivotal role of protein acetylation in linking glucose and fatty acid metabolism to β-cell function. Cell Death Dis. 2019, 10, 66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Alleti, R.; Oh, W.S.; Perambuduru, M.; Afrasiabi, Z.; Sinn, E.; Reddy, V.P. Gadolinium triflate immobilized in imidazolium based ionic liquids: A recyclable catalyst and green solvent for acetylation of alcohols and amines. Green Chem. 2005, 7, 203–206. [Google Scholar] [CrossRef]
  15. El Seoud, O.A.; Koschella, A.; Fidale, L.C.; Dorn, S.; Heinze, T. Applications of Ionic Liquids in Carbohydrate Chemistry: A Window of Opportunities. Biomacromolecules 2007, 8, 2629–2647. [Google Scholar] [CrossRef]
  16. Bartoli, G.; Dalpozzo, R.; De Nino, A.; Maiuolo, L.; Nardi, M.; Procopio, A.; Tagarelli, A. Per-O-acetylation of sugars catalyzed by Ce(OTf)3. Green Chem. 2004, 6, 191–192. [Google Scholar] [CrossRef]
  17. Deka, N.; Mariotte, A.-M.; Boumendjel, A. Microwave mediated solvent-free acetylation of deactivated and hindered phenols. Green Chem. 2001, 3, 263–264. [Google Scholar] [CrossRef]
  18. Snodin, D.J. Genotoxic Impurities: From Structural Alerts to Qualification. Org. Proc. Res. Dev. 2010, 14, 960–976. [Google Scholar] [CrossRef]
  19. Plutschack, M.B.; Pieber, B.; Gilmore, K.; Seeberger, P.H. The Hitchhiker’s Guide to Flow Chemistry. Chem. Rev. 2017, 117, 11796–11893. [Google Scholar] [CrossRef]
  20. Vukelić, S.; Koksch, B.; Seeberger, P.H.; Gilmore, K. A Sustainable, Semi-Continuous Flow Synthesis of Hydantoins. Chem. Eur. J. 2016, 22, 13451–13454. [Google Scholar] [CrossRef]
  21. Ushakov, D.B.; Gilmore, K.; Kopetzki, D.; McQuade, D.T.; Seeberger, P.H. Continuous-Flow Oxidative Cyanation of Primary and Secondary Amines Using Singlet Oxygen. Angew. Chem. Int. Ed. 2014, 45, 557. [Google Scholar] [CrossRef] [PubMed]
  22. McQuade, D.T.; Seeberger, P.H. Applying Flow Chemistry: Methods, Materials, and Multistep Synthesis. J. Org. Chem. 2013, 78, 6384–6389. [Google Scholar] [CrossRef]
  23. Watts, K.; Baker, A.; Wirth, T. Electrochemical Synthesis in Microreactors. J. Flow Chem. 2015, 4, 2–11. [Google Scholar] [CrossRef]
  24. Wiles, C.; Watts, P. Continuous process technology: A tool for sustainable production. Green Chem. 2014, 16, 55–62. [Google Scholar] [CrossRef]
  25. Wiles, C.; Watts, P. Continuous flow reactors: A perspective. Green Chem. 2012, 14, 38–54. [Google Scholar] [CrossRef]
  26. Wiles, C.; Watts, P. Solid-Supported Gallium Triflate: An Efficient Catalyst for the Three-Component Ketonic Strecker Reaction. ChemSusChem 2012, 5, 332–338. [Google Scholar] [CrossRef]
  27. Osorio-Planes, L.; Rodriguez-Escrich, C.; Pericàs, M.A. Removing the Superfluous: A Supported Squaramide Catalyst with a Minimalistic Linker Applied to the Enantioselective Flow Synthesis of Pyranonaphthoquinones. Catal. Sci. Technol. 2016, 6, 4686–4689. [Google Scholar] [CrossRef]
  28. Llanes, P.; Rodríguez-Escrich, C.; Sayalero, S.; Pericàs, M.A. Organocatalytic Enantioselective Continuous-Flow Cyclopropanation. Org. Lett. 2016, 18, 6292–6295. [Google Scholar] [CrossRef]
  29. Izquierdo, J.; Pericàs, M.A. A Recyclable, Immobilized Analogue of Benzotetramisole for Catalytic Enantioselective Domino Michael Addition/Cyclization Reactions in Batch and Flow. ACS Catal. 2016, 6, 348–356. [Google Scholar] [CrossRef] [Green Version]
  30. Pascanu, V.; Hansen, P.R.; Gómez, A.B.; Ayats, C.; Platero-Prats, A.E.; Johansson, M.J.; Pericàs, M.À.; Martín-Matute, B. Highly Functionalized Biaryls via Suzuki–Miyaura Cross-Coupling Catalyzed by Pd@MOF under Batch and Continuous Flow Regimes. ChemSusChem 2015, 8, 123–130. [Google Scholar] [CrossRef]
  31. Martin-Rapun, R.; Sayalero, S.; Pericàs, M.A. Asymmetric anti-Mannich Reactions in Continuous Flow. Green Chem. 2013, 15, 3295–3301. [Google Scholar] [CrossRef]
  32. Bottecchia, C.; Rubens, M.; Gunnoo, S.B.; Hessel, V.; Madder, A.; Noel, T. Visible-Light-Mediated Selective Arylation of Cysteine in Batch and Flow. Angew. Chem. Int. Ed. 2017, 56, 12702–12707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Su, Y.; Kuijpers, K.P.L.; König, N.; Shang, M.; Hessel, V.; Noël, T. A Mechanistic Investigation of the Visible-Light Photocatalytic Trifluoromethylation of Heterocycles Using CF3I in Flow. Chem. Eur. J. 2016, 22, 12295–12300. [Google Scholar] [CrossRef] [PubMed]
  34. Su, Y.; Kuijpers, K.; Hessel, V.; Noël, T. A Convenient Numbering-Up Strategy for the Scale-Up of Gas-Liquid Photoredox Catalysis in Flow. React. Chem. Eng. 2016, 1, 73–81. [Google Scholar] [CrossRef]
  35. Straathof, N.J.W.; Cramer, S.E.; Hessel, V.; Noël, T. Practical Photocatalytic Trifluoromethylation and Hydrotrifluoromethylation of Styrenes in Batch and Flow. Angew. Chem. Int. Ed. 2016, 55, 15549–15553. [Google Scholar] [CrossRef] [Green Version]
  36. Seo, H.; Katcher, M.H.; Jamison, T.F. Photoredox Activation of Carbon Dioxide for Amino Acid Synthesis in Continuous Flow. Nat. Chem. 2016, 9, 453–456. [Google Scholar] [CrossRef] [Green Version]
  37. McTeague, T.A.; Jamison, T.F. Photoredox Activation of SF6 for Fluorination. Angew. Chem. Int. Ed. 2016, 55, 15072–15075. [Google Scholar] [CrossRef]
  38. Adamo, A.; Beingessner, R.L.; Behnam, M.; Chen, J.; Jamison, T.F.; Jensen, K.F.; Monbaliu, J.C.M.; Myerson, A.S.; Revalor, E.M.; Snead, D.R. On-Demand Continuous-Flow Production of Pharmaceuticals in a Compact, Reconfigurable System. Science 2016, 352, 61–67. [Google Scholar] [CrossRef] [Green Version]
  39. Snead, D.R.; Jamison, T.F. A Three-Minute Synthesis and Purification of Ibuprofen: Pushing the Limits of Continuous-Flow Processing. Angew. Chem. Int. Ed. 2015, 54, 983–987. [Google Scholar] [CrossRef]
  40. Wu, J.; Yang, X.; He, Z.; Mao, X.; Hatton, T.A.; Jamison, T.F. Continuous Flow Synthesis of Ketones from Carbon Dioxide and Organolithium or Grignard Reagents. Angew. Chem. Int. Ed. 2014, 53, 8416–8420. [Google Scholar] [CrossRef]
  41. Rogers, L.; Jensen, K.F. Continuous manufacturing – the Green Chemistry promise? Green Chem. 2019, 21, 3481–3498. [Google Scholar] [CrossRef] [Green Version]
  42. Giri, G.; Yang, L.; Mo, Y.; Jensen, K.F. Adding Crystals to Minimize Clogging in Continuous Flow Synthesis. Cryst. Growth Des. 2019, 19, 98–105. [Google Scholar] [CrossRef]
  43. Coley, C.W.; Thomas, D.A.; Lummiss, J.A.M.; Jaworski, J.N.; Breen, C.P.; Schultz, V.; Hart, T.; Fishman, J.S.; Rogers, L.; Gao, H.; et al. A robotic platform for flow synthesis of organic compounds informed by AI planning. Science 2019, 365. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, P.; Weeranoppanant, N.; Thomas, D.A.; Tahara, K.; Stelzer, T.; Russell, M.G.; O’Mahony, M.; Myerson, A.S.; Lin, H.; Kelly, L.P.; et al. Advanced Continuous Flow Platform for On-Demand Pharmaceutical Manufacturing. Chem. Eur. J. 2018, 24, 2776–2784. [Google Scholar] [CrossRef]
  45. Bédard, A.-C.; Adamo, A.; Aroh, K.C.; Russell, M.G.; Bedermann, A.A.; Torosian, J.; Yue, B.; Jensen, K.F.; Jamison, T.F. Reconfigurable system for automated optimization of diverse chemical reactions. Science 2018, 361, 1220–1225. [Google Scholar] [CrossRef] [Green Version]
  46. Jensen, K.F. Flow chemistry. Microreaction technology comes of age. AIChE J. 2017, 63, 858–869. [Google Scholar] [CrossRef]
  47. Ollivier, N.; Toupy, T.; Hartkoorn, R.C.; Desmet, R.; Monbaliu, J.-C.M.; Melnyk, O. Accelerated microfluidic native chemical ligation at difficult amino acids toward cyclic peptides. Nat. Commun. 2018, 9, 2847. [Google Scholar] [CrossRef] [Green Version]
  48. Kreituss, I.; Bode, J.W. Flow chemistry and polymer-supported pseudoenantiomeric acylating agents enable parallel kinetic resolution of chiral saturated N-heterocycles. Nat. Chem. 2016, 9, 446–452. [Google Scholar] [CrossRef] [Green Version]
  49. Berton, M.; Huck, L.; Alcázar, J. On-demand synthesis of organozinc halides under continuous flow conditions. Nat. Protoc. 2018, 13, 324–334. [Google Scholar] [CrossRef]
  50. Elvira, K.S.; Solvas, X.C.I.; Wootton, R.C.R.; de Mello, A.J. The past, present and potential for microfluidic reactor technology in chemical synthesis. Nat. Chem. 2013, 5, 905–915. [Google Scholar] [CrossRef]
  51. Tadele, K.; Verma, S.; Nadagouda, M.N.; Gonzalez, M.A.; Varma, R.S. A rapid flow strategy for the oxidative cyanation of secondary and tertiary amines via C-H activation. Sci. Rep. 2017, 7, 16311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Fuse, S.; Mifune, Y.; Nakamura, H.; Tanaka, H. Total synthesis of feglymycin based on a linear/convergent hybrid approach using micro-flow amide bond formation. Nat. Commun. 2016, 7, 13491. [Google Scholar] [CrossRef] [PubMed]
  53. Kirschneck, D. Strategies for Using Microreactors and Flow Chemistry: Drivers and Tools. Chem. Eng. Tech. 2013, 36, 1061–1066. [Google Scholar] [CrossRef]
  54. Fanelli, F.; Parisi, G.; Degennaro, L.; Luisi, R. Contribution of microreactor technology and flow chemistry to the development of green and sustainable synthesis. Beilstein J. Org. Chem. 2017, 13, 520–542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. De Angelis, S.; Celestini, P.; Purgatorio, R.; Degennaro, L.; Rebuzzini, G.; Luisi, R.; Carlucci, C. Development of a continuous flow synthesis of propranolol: Tackling a competitive side reaction. J. Flow Chem. 2019, 9, 231–236. [Google Scholar] [CrossRef]
  56. De Angelis, S.; Hone, C.; Degennaro, L.; Celestini, P.; Luisi, R.; Kappe, C.O. Sequential α-lithiation and aerobic oxidation of an arylacetic acid-continuous-flow synthesis of cyclopentyl mandelic acid. J. Flow Chem. 2018, 8, 109–116. [Google Scholar] [CrossRef] [Green Version]
  57. Movsisyan, M.; Delbeke, E.I.P.; Berton, J.K.E.T.; Battilocchio, C.; Ley, S.V.; Stevens, C.V. Taming hazardous chemistry by continuous flow technology. Chem. Soc. Rev. 2016, 45, 4892–4928. [Google Scholar] [CrossRef]
  58. Mellmer, M.A.; Sanpitakseree, C.; Demir, B.; Bai, P.; Ma, K.; Neurock, M.; Dumesic, J.A. Solvent-enabled control of reactivity for liquid-phase reactions of biomass-derived compounds. Nat. Catal. 2018, 1, 199–207. [Google Scholar] [CrossRef]
  59. Xiong, B.; Wang, G.; Xiong, T.; Wan, L.; Zhou, C.; Liu, Y.; Zhang, P.; Yang, C.; Tang, K. Direct synthesis of N-arylamides via the coupling of aryl diazonium tetrafluoroborates and nitriles under transition-metal-free conditions. Tetrahedron Lett. 2018, 59, 3139–3142. [Google Scholar] [CrossRef]
  60. Lee, Y.M.; Moon, M.E.; Vajpayee, V.; Filimonov, V.D.; Chi, K.-W. Efficient and economic halogenation of aryl amines via arenediazonium tosylate salts. Tetrahedron 2010, 66, 7418–7422. [Google Scholar] [CrossRef]
  61. Beller, M. First Amidocarbonylation with Nitriles for the Synthesis of N-Acyl Amino Acids. Synlett 1999, 1999, 108–110. [Google Scholar] [CrossRef]
  62. Enders, D.; Shilvock, J. Some recent applications of α -amino nitrile chemistry. Chem. Soc. Rev. 2000, 29, 359–373. [Google Scholar] [CrossRef]
  63. Jiang, T.-S.; Wang, G.-W. Synthesis of 3-Acylindoles by Palladium-Catalyzed Acylation of Free (N–H) Indoles with Nitriles. Org. Lett. 2013, 15, 788–791. [Google Scholar] [CrossRef] [PubMed]
  64. Jang, W.B.; Shin, W.S.; Hong, J.E.; Lee, S.Y.; Oh, D.Y. Synthesis of Ketones with Alkyl Phosphonates and Nitriles as Acyl Cation Equivalent: Application of Dephosphonylation Reaction of β-Functionalized Phosphonate with Hydride. Synth. Commun. 1997, 27, 3333–3339. [Google Scholar] [CrossRef]
  65. Saikia, U.P.; Hussain, F.L.; Suri, M.; Pahari, P. Selective N-acetylation of aromatic amines using acetonitrile as acylating agent. Tetrahedron Lett. 2016, 57, 1158–1160. [Google Scholar] [CrossRef]
  66. Brahmayya, M.; Suen, S.-Y.; Dai, S.A. Sulfonated graphene oxide-catalyzed N-acetylation of amines with acetonitrile under sonication. J. Taiwan Inst. Chem. Eng. 2018, 83, 174–183. [Google Scholar] [CrossRef]
  67. Molnár, Á.; Papp, A. Catalyst recycling—A survey of recent progress and current status. Coord. Chem. Rev. 2017, 349, 1–65. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds 128 are available from the authors.
Figure 1. Schematic representation of the reactor used in the study.
Figure 1. Schematic representation of the reactor used in the study.
Molecules 25 01985 g001
Figure 2. The effect of temperature (a), pressure (b), flow rate (c) and concentration (d) on the reaction conversion catalyzed by Al2O3. The effect of the pressure was measured at room temperature and the effect of temperature was determined at 50 bar, while the effect of the flow rate and concentration was analyzed at the optimized conditions.
Figure 2. The effect of temperature (a), pressure (b), flow rate (c) and concentration (d) on the reaction conversion catalyzed by Al2O3. The effect of the pressure was measured at room temperature and the effect of temperature was determined at 50 bar, while the effect of the flow rate and concentration was analyzed at the optimized conditions.
Molecules 25 01985 g002
Figure 3. Robustness of the acetylation investigated in the reaction of benzylamine. The same reaction was repeated 10 times on the same catalyst.
Figure 3. Robustness of the acetylation investigated in the reaction of benzylamine. The same reaction was repeated 10 times on the same catalyst.
Molecules 25 01985 g003
Scheme 1. The suspected reaction mechanism of acetylation.
Scheme 1. The suspected reaction mechanism of acetylation.
Molecules 25 01985 sch001
Table 1. Lewis acids screen in continuous-flow (CF) acetylation.
Table 1. Lewis acids screen in continuous-flow (CF) acetylation.
Molecules 25 01985 i001
EntryLewis AcidsSolventConversion (%)
1Fe2O3Acetonitrile0
2Boric acidAcetonitrile3
3AlCl3Acetonitrile19
4Al2O3Acetonitrile64
Conditions: 150 °C, 50 bar, 0.1 mL min−1, 27 min residence time in flow synthesis.
Table 2. Results of the acetylation of various amines obtained at the optimal conditions: 200 °C, 50 bar, 0.1 mL min−1 flow rate, 27 min residence time.
Table 2. Results of the acetylation of various amines obtained at the optimal conditions: 200 °C, 50 bar, 0.1 mL min−1 flow rate, 27 min residence time.
EntrySubstrateProductYield (%)Space Time Yield
(mol kg−1 h−1)
1 Molecules 25 01985 i002
1
Molecules 25 01985 i003
2
> 991.5
2 Molecules 25 01985 i004
3
Molecules 25 01985 i005
4
931.395
3 Molecules 25 01985 i006
5
Molecules 25 01985 i007
6
510.765
4 Molecules 25 01985 i008
7
Molecules 25 01985 i009
8
> 991.5
5 Molecules 25 01985 i010
9
Molecules 25 01985 i011
10
> 991.5
6 Molecules 25 01985 i012
11
Molecules 25 01985 i013
12
951.425
7 Molecules 25 01985 i014
13
Molecules 25 01985 i015
14
00
8 Molecules 25 01985 i016
15
Molecules 25 01985 i017
16
00
9 Molecules 25 01985 i018
17
Molecules 25 01985 i019
18
00
10 Molecules 25 01985 i020
19
Molecules 25 01985 i021
20
> 991.5
11 Molecules 25 01985 i022
21
Molecules 25 01985 i023
22
> 991.5
12 Molecules 25 01985 i024
23
Molecules 25 01985 i025
24
> 991.5
13 Molecules 25 01985 i026
25
Molecules 25 01985 i027
26
> 991.5
14 Molecules 25 01985 i028
27
Molecules 25 01985 i029
28
> 991.5

Share and Cite

MDPI and ACS Style

Orsy, G.; Fülöp, F.; Mándity, I.M. N-Acetylation of Amines in Continuous-Flow with Acetonitrile—No Need for Hazardous and Toxic Carboxylic Acid Derivatives. Molecules 2020, 25, 1985. https://doi.org/10.3390/molecules25081985

AMA Style

Orsy G, Fülöp F, Mándity IM. N-Acetylation of Amines in Continuous-Flow with Acetonitrile—No Need for Hazardous and Toxic Carboxylic Acid Derivatives. Molecules. 2020; 25(8):1985. https://doi.org/10.3390/molecules25081985

Chicago/Turabian Style

Orsy, György, Ferenc Fülöp, and István M. Mándity. 2020. "N-Acetylation of Amines in Continuous-Flow with Acetonitrile—No Need for Hazardous and Toxic Carboxylic Acid Derivatives" Molecules 25, no. 8: 1985. https://doi.org/10.3390/molecules25081985

Article Metrics

Back to TopTop