Next Article in Journal
The Assessment of Meloxicam Phototoxicity in Human Normal Skin Cells: In Vitro Studies on Dermal Fibroblasts and Epidermal Melanocytes
Next Article in Special Issue
One-Pot Solvent-Involved Synthesis of 5-O-Substituted 5H-Chromeno[2,3-b]pyridines
Previous Article in Journal
Simultaneous Estimation of Escitalopram and Clonazepam in Tablet Dosage Forms Using HPLC-DAD Method and Optimization of Chromatographic Conditions by Box-Behnken Design
Previous Article in Special Issue
Co-N-Si/AC Catalyst for Aerobic Oxidation of Benzyl Alcohols to Esters under Mild Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Performance of Immobilized Sulfuric Acid on Silica Gel for N-Formylation of Amines with Triethyl Orthoformate

by
Sodeeq Aderotimi Salami
1,*,
Xavier Siwe-Noundou
2,* and
Rui W. M. Krause
1,*
1
Department of Chemistry, Rhodes University, Grahamstown 6140, South Africa
2
Department of Pharmaceutical Sciences, School of Pharmacy, Sefako Makgatho Health Sciences University, P.O Box 218, Pretoria 0204, South Africa
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(13), 4213; https://doi.org/10.3390/molecules27134213
Submission received: 25 April 2022 / Revised: 13 June 2022 / Accepted: 16 June 2022 / Published: 30 June 2022
(This article belongs to the Special Issue Green and Highly Efficient One-Pot Synthesis and Catalysis)

Abstract

:
In the search for convenient, green, and practical catalytic methods for the current interest in organic synthesis, a simple, green, and highly efficient protocol for N-formylation of various amines was carried out in the presence of immobilized sulfuric acid on silica gel (H2SO4–SiO2). All reactions were performed in refluxing triethyl orthoformate (65 °C). The product formamides were obtained with high-to-excellent yields within 4 min to 2 h. The current approach is advantageous, due to its short reaction time and high yields. The catalyst is recyclable with no significant loss in catalytic efficiency.

1. Introduction

A fascinating trend in the synthesis of widely used organic molecules is the focus on green chemistry, including efficient reactions and the use of ecologically friendly reagents [1]. The use of silica gel as an effective catalyst in chemical processes has attracted much attention in recent years. The formylation of amines is a crucial process in organic chemistry, owing to the widespread application of N-formyl amine derivatives in industry and in biologically active compounds, such as fluoroquinolones, substituted imidazoles, 1,2-dihydroquinolines, and nitrogen-bridged heterocycles, among others [2]. N-formyl amine derivatives have also been used as reagents in Vilsmeier formylation reactions as amino acid-protecting groups [3] and in the synthesis of several other important derivatives, such as formamidines [4], isocyanates [5], and nitriles [6] (Figure 1).
Despite the fact that there are a variety of reagents for N-formylation of amines, the synthesis of formamides utilizing triethyl orthoformate as a formylating agent is still popular [1]. The reaction of ethyl orthoformate with aniline to afford N,N′-diphenylformamidine was initially reported in 1869 by Wichelhaus [7]. Subsequently, Claisen synthesized ethyl N-phenylformimidate in low yields from the same reactants, but under slightly different experimental conditions [8]. Swaringen and colleagues went on to show that the reaction of N-alkylanilines with orthoformates in the absence of a catalyst or with hydrochloric/acetic acid produced orthoamides in low yields [8]. These few examples demonstrate one of the major drawbacks of this system, namely, the low yield. Meanwhile, when p-toluenesulfonic acid was employed as a catalyst, high yields of N-alkylformanilides and N,N-dialkylanilines were generated, but the reactions still often required high temperatures and prolonged reaction times. For example, Swaringen and co-workers demonstrated the synthesis of N-ethyl formamides from the reaction of amines with triethyl orthoformate in the presence of H2SO4, but under severe conditions (temperature above 140 °C) [9].
Various other formylating agents have been reported, including chloral [10], acetic formic anhydride [11], formic acid [12], ammonium formate [13], formate esters [14], polymer-supported formate [15], ethyl formate [16], triethyl orthoformate [1,2], aldehydes and methanol [17], carbon monoxide [18], and carbon dioxide [19]. However, these also tend to suffer from similar problems of long reaction times (hours to days), variable or low yields, and harsh conditions (or expensive catalyst systems).
Several catalysts have been employed for the formylation of amines, including silica-supported sulfuric acid [20], H2SO4/NaHSO4-activated charcoal [21], K-F alumina [22], Amberlite IR 120 [23], ZnO [24], nano-CeO2 [25], nano-MgO [26], natrolite zeolite [27], indium metal [28], sulfated titania [29], and sulfated tungstate [30], among others (Table 1).
In the absence of a catalyst or promoter, N-formylation of amines is a sluggish reaction that usually requires unique reaction conditions or long time frames for completion [25]. However, some of these methods have quite a number of limitations, including harsh reaction conditions, the need for expensive metal catalysts or organocatalysts, and long reaction time frames. Thus, for organic transformations, the development of a safe, benign, environmentally friendly, high-yield, quick-reaction, and recyclable catalyst for N-formylation of amines remains extremely desirable [3].
In the last few years, H2SO4–SiO2 (Table 2) has demonstrated significant promise as a cost-effective and easily retrievable solid catalyst for driving a variety of essential organic reactions in solvent-free environments. H2SO4–SiO2 is appealing for industrial usage because of its high catalytic activity, operational simplicity, and recyclability. There are two types of functional groups on the silica surface: siloxane (Si–O–Si) and silanol (Si–OH). Thus, silica gel modification can occur through the reaction of a specific molecule with either the siloxane (nucleophilic substitution at the Si) or silanol (direct reaction with the hydroxyl group) functions, though it is widely accepted that the reaction with the silanol function is the most common modification pathway (Figure 2) [47,48]. The notion of employing H2SO4–SiO2 as a transamidation catalyst was inspired by Rasheed et al. [20]. We became interested in employing the same catalyst to build a generic formylation with triethyl orthoformate. To the best of our knowledge, no reports of H2SO4–SiO2-catalyzed formylation with triethyl orthoformate have been published, and so for the first time, we present findings in this regard.

2. Results and Discussion

Initially, the reaction of aniline with triethyl orthoformate was chosen as the model reaction (Figure 3). During the optimization of reaction parameters, it was observed that aniline reacted smoothly with triethyl orthoformate, providing the desired product with a good yield (96%) within a short period of time (Table 3).
In order to generalize the protocol for the formylation of sterically hindered amines, the reaction was optimized with respect to temperature and molar ratio. The temperature was raised to 65 °C and was observed to be quite sufficient to carry out the reaction with an optimum yield of the desired product (Table 3). It was observed that the need for an excess of triethyl orthoformate was no longer required, as a 1:3 molar ratio of amine to triethyl orthoformate was sufficient to yield the desired product (Table 3, entry 3).
We next explored the impact of immobilized sulfuric acid on silica gel stoichiometry on the outcome of the reaction (Table 4). We observed that excess H2SO4–SiO2 was not beneficial for faster conversion. Conversely, a lower amount of H2SO4–SiO2 led to substantially slower conversion. The background reaction (used as a model) was also measured in the absence of H2SO4–SiO2, confirming its vital role.
In general, the reaction proceeded efficiently, with various amines reacting with triethyl orthoformate to produce the corresponding N-formylated product with good-to-excellent yield within a very short time. Aliphatic and aromatic primary amines underwent smooth N-formylation and gave the product in 70–96% yields (Table 5).
Aniline with electron-donating groups provided an excellent yield of 65–96% with triethyl orthoformate. The halogen (F, Cl, Br, I)-containing anilines provided good yields, ranging from 73% to 96%, of corresponding products. Similarly, electron-withdrawing groups were found to react smoothly under the optimized reaction conditions and demonstrate good yields of desired products (85–96%). Generally, under these optimized reaction conditions, various functional groups were tolerated. However, finding a general method for generating amide bonds will surely benefit the drug discovery process. In general, the formylation of aryl/heteroaryl amines (electron-neutral, -rich, -deficient), aliphatic, and cyclic secondary amines afforded the formylation products in excellent yields (70–96%). Interestingly, sterically hindered aryl amines, such as products 6, 7, 10, 11, 16, 17, and 3338, were found to react smoothly under the optimized reaction conditions, demonstrating good yields of desired products. Less reactive hetero aromatics, such as 4251 and 56, produced the product with a surprisingly high yield (77–90%) and a longer reaction time (35–60 min). When secondary amines 5254 were employed, the reaction was somehow slow, providing a good yield of products in 1 h (Table 5). NMR spectral data of all synthesized compounds are available in the Supplementary Materials (S1–S56).

3. Reusability of Catalyst

The reusability of the catalytic system was explored. The catalyst was separated by simple filtration and washed with ethyl acetate after the reaction was completed, and it was reused for two consecutive cycles within the same time frame (4 min), with a slight decrease in catalytic activity (9–13%) (Table 6).
In order to demonstrate the efficiency and versatility of the H2SO4–SiO2 system, we compared the result of N-formylation of aniline with other protocols that have been published based on reaction times and yields (Table 7). The results showed that the other approaches required longer reaction times for efficient conversion than for the present protocol. Therefore, on this basis, the present protocol is more efficient or comparable with other methodologies.
Even though we have yet to prove the mechanism of our reaction in an experimental manner, Figure 4 suggests a possible explanation. The first step is the activation of the electrophilic carbon of triethyl orthoformate by the sulfonic group of H2SO4–SiO2, which led to the formation of a cationic intermediate. The cationic intermediate reacted with amine nucleophiles, which, on further elimination of ethanol, furnished the desired formylated product.
While 1,8-difformamido-naphthalein (38) and 3-formamido-1,2,4-triazole-5-thiol (53) are new derivatives and were characterized by one- and two-dimensional NMR analysis and high-resolution mass spectroscopy, all other products are known compounds and were identified by melting point, IR, 1H NMR, and 13C NMR spectroscopy. The synthesis of formamides was confirmed by IR spectra, which revealed two distinct absorption bands between 3300 and 3400 cm−1 (secondary NH) and 1640 and 1680 cm−1 (N-formyl, C=O).
Furthermore, formamide molecules have both a conformational stereogenic axis and a configurational stereogenic centre. These molecules take on two distinct syn and anti-conformational diastereomers as a result of restricted rotation around the Ar–N bond [50]. The 1H and 13C NMR spectra of most of the synthesized formamides at 25 °C were consistent with the presence of two rotamers. Only one rotamer was observed for the compounds 8, 14, 27, 45 and 46.
During the purification of compounds 12 and 35, two products appeared as partially separated spots on thin-layer chromatography (TLC) plates. Using normal silica gel chromatography, these compounds were identified as A and B rotamer pairs. After purifying compounds 12 and 35, pure rotamers 12A and 35A were isolated (Figure 5). 12A and 35A were the only pure isomers that could be isolated, while 12B and 35B were always contaminated to some degree by 12A and 35A, respectively. The fact that we were able to isolate rotamers A and B at room temperature and characterize them using basic spectroscopic techniques astounded us. This occurrence may be viewed as a specific form of atropisomerism, because atropisomers are stereoisomers with restricted rotation around a single bond where the rotational barrier is high enough to allow isolation of the isomeric species [51].

4. Materials and Methods

A PerkinElmer Spectrum 100 FT-IR Spectrometer (Valencia, CA, USA) was used for the FT-IR analysis. The IR spectra were obtained by the attenuated total reflection (ATR) method. For each experiment, 16 scans were performed in the frequency range from 650 to 4000 cm−1. Melting points of all the compounds were determined using a Koffler hot-stage apparatus and were uncorrected. NMR spectra were recorded on a Bruker Advance III 400 spectrometer (Rheinstetten, Germany) using CDCl3 or DMSO-d6 as a solvent with tetramethyl silane used as internal standard. LC-MS/MS data were recorded on a Bruker Compact quadrupole time of flight (QToF) mass spectrometer (Bremen, Germany). Raw mass spectrometry data were processed using MZmine software (version 2.38) (San Diego, CA, USA). Solvents and chemicals used were of analytical grade, purchased from Sigma Aldrich (St. Louis, MO, USA) and used without further purification. The purity determination of the starting materials and reaction monitoring were performed by thin-layer chromatography (TLC) on Merck silica gel G F254plates (Duren, Germany).

4.1. Preparation of Sulfuric Acid Adsorbed on Silica Gel (H2SO4–SiO2)

The preparation of H2SO4–SiO2 was carried out by following the reported procedure [52]. To a suspension of silica gel (29.5 g, 230–400 mesh size) in EtOAc (60 mL), H2SO4 (1.5 g, 15.5 mmol, 0.8 mL of a 98% aq. solution of H2SO4) was added and the mixture was stirred magnetically for 30 min at room temperature. EtOAc was removed under reduced pressure (rotary evaporator) and the residue was heated at 100 °C for 72 h under vacuum to afford H2SO4–SiO2 as a free-flowing powder.

4.2. A General Procedure for N-Formylation of Amines with Triethyl Orthoformate Promoted by Immobilized H2SO4 on Silica Gel

To a mixture of aniline (0.548 mL, 6 mmol) and triethyl orthoformate (24 mmol), the immobilized H2SO4 on silica gel (1.2 g) was then added and the reaction mixture was stirred under reflux conditions (65 °C). Progress of the reaction was monitored by TLC. After completion of the reaction, the mixture was diluted with EtOAc (20 mL), filtered, water (30 mL) was added, the solution was extracted with EtOAc, and the combined organic layers were dried over anhydrous Na2SO4 and concentrated. The residue was subjected to column chromatography and eluted with (EtOAc–Pet Ether (3:1)) to afford the product in high yields.

5. Conclusions

We have developed a simple, green, and highly efficient protocol for N-formylation of various amines in the presence of immobilized sulfuric acid on silica gel, with excellent yields and remarkably simple and environmentally benign processes. The approach is compatible with a wide range of aromatic, heteroaromatic, aliphatic, and cyclic/acyclic primary or secondary amines. The H2SO4–SiO2 catalytic system described here is a good complement to previously reported protocols, due to its ease of manipulation, low cost, and benign nature. We are optimistic that, with this approach, we will be able to develop the biologically relevant heterocyclic ring system more efficiently. This protocol is generic, and it will undoubtedly offer value to the growing area of organic synthesis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27134213/s1: Figure S1–S56: NMR spectral data of synthesized compounds. References [53,54,55,56,57,58] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, R.W.M.K. and S.A.S.; methodology, S.A.S.; writing—original draft preparation, S.A.S.; writing—review and editing, X.S.-N.; supervision, R.W.M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Research Foundation of South Africa (Grant No. 116109).

Data Availability Statement

Original data from experiments are available from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Habibi, D.; Sahebekhtiari, H.; Nasrollahzadeh, M.; Taghipour, A. A Very Simple, Highly Efficient and Catalyst-free Procedure for the N-Formylation of Amines Using Triethyl orthoformate in Water Under Ultrasound-irradiation. Lett. Org. Chem. 2013, 10, 209–212. [Google Scholar] [CrossRef]
  2. Kaboudin, B.; Khodamorady, M. Organic reactions in water: A practical and convenient method for the N-formylation of amines in water. Synlett 2010, 19, 2905–2907. [Google Scholar] [CrossRef]
  3. Khatri, C.K.; Chaturbhuj, G.U. Sulfated polyborate-catalyzed N-formylation of amines: A rapid, green and efficient protocol. J. Iran. Chem. Soc. 2017, 14, 2513–2519. [Google Scholar] [CrossRef]
  4. Han, Y.; Cai, L. An efficient and convenient synthesis of formamidines. Tetrahedron Lett. 1997, 38, 5423–5426. [Google Scholar] [CrossRef]
  5. Gould-Fogerite, S.; Mannino, R.J. Protein or peptide-cochleate vaccines and methods of immunizing using the same. U.S. Patent 5,643,574, 1 July 1997. [Google Scholar]
  6. Pravin, P.; Maryam, A.M.; Alexander, D. Isocyanide 2.0. Green Chem. 2020, 22, 6902–6911. [Google Scholar] [CrossRef]
  7. Roberts, R.M.; Vogt, P.J. Ortho esters, imidic esters and amidines: N-alkylformanilides from alkyl orthoformates and primary aromatic amines; Rearrangement of alkyl N-arylformimidates. J. Am. Chem. Soc. 1956, 78, 4778–4782. [Google Scholar] [CrossRef]
  8. de la Mare, P.B.D. Kinetics of thermal addition of halogens to olefinic compounds. Q. Rev. Chem. Soc. 1949, 3, 126–145. [Google Scholar] [CrossRef]
  9. Swaringen, R.A.; Eaddy, J.F.; Henderson, T.R. Reaction of Ortho Esters with Secondary Amines. J. Org. Chem. 1980, 45, 3986–3989. [Google Scholar] [CrossRef]
  10. Blicke, F.F.; Lu, C.-J. Formylation of Amines with Chloral and Reduction of the N-Formyl Derivatives with Lithium Aluminum Hydride. J. Am. Chem. Soc. 1952, 74, 3933–3934. [Google Scholar] [CrossRef]
  11. Kiho, T.; Yoshida, I.; Katsuragawa, M.; Sakushima, M.; Usui, S.; Ukai, S. Polysaccharides in Fungi. XXXIV. A Polysaccharide from the Fruiting Bodies of Amanita muscaria and the Antitumor Activity of Its Carboxymethylated Product. Biol. Pharm. Bull. 1994, 17, 1460–1462. [Google Scholar] [CrossRef] [Green Version]
  12. Jung, S.H.; Ahn, J.H.; Park, S.K.; Choi, J.K. A practical and convenient procedure for the N-formylation of amines using formic acid. Bull. Korean Chem. Soc. 2002, 23, 149–150. [Google Scholar] [CrossRef] [Green Version]
  13. Ganapati Reddy, P.; Kishore Kumar, G.D.; Baskaran, S. A convenient method for the N-formylation of secondary amines and anilines using ammonium formate. Tetrahedron Lett. 2000, 41, 9149–9151. [Google Scholar] [CrossRef]
  14. Rupesh Patre, E.; Sanjib Mal, A.; Pankaj, R.; Nilkanth, R.; Sujit Ghorai, K.; Sudhindra Deshpande, H.; Myriem Qacemi, E.I.; Smejkal, T.; Pal, S.; Manjunath, B.N. First report on bio-catalytic N-formylation of amines using ethyl formate. Chem. Commun. 2017, 53, 2382–2385. [Google Scholar] [CrossRef] [PubMed]
  15. Desai, B.; Danks, T.N.; Wagner, G. Thermal and microwave-assisted N-formylation using solid-supported reagents. Tetrahedron Lett. 2005, 46, 955–957. [Google Scholar] [CrossRef]
  16. Dhake, K.P.; Tambade, P.J.; Singhal, R.S.; Bhanage, B.M. An efficient, catalyst- and solvent-free N-formylation of aromatic and aliphatic amines. Green Chem. Lett. Rev. 2011, 4, 151–157. [Google Scholar] [CrossRef] [Green Version]
  17. Das, B.; Krishnaiah, M.; Balasubramanyam, P.; Veeranjaneyulu, B.; Nandan Kumar, D. A remarkably simple N-formylation of anilines using polyethylene glycol. Tetrahedron Lett. 2008, 49, 2225–2227. [Google Scholar] [CrossRef]
  18. Noh, H.W.; An, Y.; Lee, S.; Jung, J.; Son, S.U.; Jang, H.Y. Metal-free Carbon Monoxide (CO) Capture and Utilization: Formylation of Amines. Adv. Synth. Catal. 2019, 361, 3068–3073. [Google Scholar] [CrossRef]
  19. Zhang, L.; Han, Z.; Zhao, X.; Wang, Z.; Ding, K. Highly efficient ruthenium-catalyzed N-formylation of amines with H2 and CO2. Angew. Chem. Int. Ed. 2015, 54, 186–6189. [Google Scholar] [CrossRef]
  20. Rasheed, S.; Rao, D.N.; Reddy, A.S.; Shankar, R.; Das, P. Sulphuric acid immobilized on silica gel (H2SO4-SiO2) as an eco-friendly catalyst for transamidation. RSC Adv. 2015, 5, 10567–10574. [Google Scholar] [CrossRef]
  21. Zeynizadeh, B.; Abdollahi, M. The immobilized NaHSO4·H2O on activated charcoal: A highly efficient promoter system for N-formylation of amines with ethyl formate. Curr. Chem. Lett. 2016, 5, 51–58. [Google Scholar] [CrossRef]
  22. Das, V.K.; Devi, R.R.; Raul, P.K.; Thakur, A.J. Nano rod-shaped and reusable basic Al2O3 catalyst for N-formylation of amines under solvent-free conditions: A novel, practical and convenient NOSE’ approach. Green Chem. 2012, 14, 847–854. [Google Scholar] [CrossRef]
  23. Madthukur Bhojegowd, M.R.; Nizam, A.; Pasha, M.A. Amberlite IR-120: A reusable catalyst for N-formylation of amines with formic acid using microwaves. Cuihua Xuebao/Chin. J. Catal. 2010, 31, 518–520. [Google Scholar] [CrossRef]
  24. Hosseini-sarvari, M.; Sharghi, H. ZnO as a New Catalyst for N-Formylation of Amines under Solvent-Free Conditions. Tetrahedron 2006, 8, 6652–6654. [Google Scholar]
  25. Zeynizadeh, B. Catalytic Performance. J. Chem. Soc. Pak. 2017, 39, 1–11. [Google Scholar]
  26. Nasrollahzadeh, M.; Motahharifar, N.; Sajjadi, M.; Aghbolagh, A.M.; Shokouhimehr, M.; Varma, R.S. Recent advances in N -formylation of amines and nitroarenes using efficient (nano)catalysts in eco-friendly media. Green Chem. 2019, 21, 5144–5167. [Google Scholar] [CrossRef]
  27. Bahari, S.; Sajadi, S.M. Natrolite zeolite: A natural and reusable catalyst for one-pot synthesis of α-aminophosphonates under solvent-free conditions. Arab. J. Chem. 2012, 10, 700–704. [Google Scholar] [CrossRef] [Green Version]
  28. Kim, J.G.; Jang, D.O. Indium-catalyzed N-formylation of amines under solvent-free conditions. Synlett 2010, 8, 1231–1234. [Google Scholar] [CrossRef]
  29. Krishnakumar, B.; Swaminathan, M. A convenient method for the N-formylation of amines at room temperature using TiO2-P25 or sulfated titania. J. Mol. Catal. A Chem. 2011, 334, 98–102. [Google Scholar] [CrossRef]
  30. Veer, S.D.; Pathare, S.P.; Akamanchi, K.G. Sulfated tungstate catalyzed hydration of alkynes. Ark. 2016, 2016, 59–66. [Google Scholar] [CrossRef] [Green Version]
  31. Thirunarayanan, G.; Muthuvel, I.; Sathiyendiran, V. Spectral LFER studies in some N-(substituted phenyl) formamides. Ann. Chem. 2015, 70, 31. [Google Scholar] [CrossRef]
  32. Kim, J.G.; Jang, D.O. Facile and highly efficient N-formylation of amines using a catalytic amount of iodine under solvent-free conditions. Synlett 2010, 14, 2093–2096. [Google Scholar] [CrossRef]
  33. Lei, M.; Ma, L.; Hu, L. A convenient one-pot synthesis of formamide derivatives using thiamine hydrochloride as a novel catalyst. Tetrahedron Lett. 2010, 51, 4186–4188. [Google Scholar] [CrossRef]
  34. Jafarzadeh, M.; Soleimani, E.; Norouzi, P.; Adnan, R.; Sepahvand, H. Preparation of trifluoroacetic acid-immobilized Fe3O4@SiO2-APTES nanocatalyst for synthesis of quinolines. J. Fluor. Chem. 2015, 178, 219–224. [Google Scholar] [CrossRef]
  35. Pathare, S.P.; Sawant, R.V.; Akamanchi, K.G. Sulfated tungstate catalyzed highly accelerated N-formylation. Tetrahedron Lett. 2012, 53, 3259–3263. [Google Scholar] [CrossRef]
  36. De Luca, L.; Giacomelli, G.; Porcheddu, A.; Salaris, M. A new, simple procedure for the synthesis of formyl amides. Synlett 2004, 14, 2570–2572. [Google Scholar] [CrossRef]
  37. Deutsch, J.; Eckelt, R.; Köckritz, A.; Martin, A. Catalytic reaction of methyl formate with amines to formamides. Tetrahedron 2009, 65, 10365–10369. [Google Scholar] [CrossRef]
  38. Baghbanian, S.M.; Farhang, M. Protic [TBD][TFA] ionic liquid as a reusable and highly efficient catalyst for N-formylation of amines using formic acid under solvent-free condition. J. Mol. Liq. 2013, 183, 45–49. [Google Scholar] [CrossRef]
  39. Chandra Shekhar, A.; Ravi Kumar, A.; Sathaiah, G.; Luke Paul, V.; Sridhar, M.; Shanthan Rao, P. Facile N-formylation of amines using Lewis acids as novel catalysts. Tetrahedron Lett. 2009, 50, 7099–7101. [Google Scholar] [CrossRef]
  40. Hong, M.; Xiao, G. Hafnium(IV) bis(perfluorooctanesulfonyl)imide complex supported on fluorous silica gel catalyzed N-formylation of amines using aqueous formic acid. J. Fluor. Chem. 2013, 146, 11–14. [Google Scholar] [CrossRef]
  41. Ourida, S.; Mark, J.B.; John, B.; James, L.; Stephen, P.M.; Pawel, P.; Robert, J.W.; Williams, J.M.J. Iridium-catalyzed formylation of amines with paraformaldehyde. Tetrahedron Lett. 2010, 51, 5804–5806. [Google Scholar] [CrossRef]
  42. Lundberg, H. Group (IV) Metal-Catalyzed Direct Amidation: Synthesis and Mechanistic Considerations. Ph.D. Thesis, University of Stockholm, Stockholm, Sweden, 2015. [Google Scholar]
  43. Ishida, T.; Haruta, M. N-formylation of amines via the aerobic oxidation of methanol over supported gold nanoparticles. ChemSusChem 2009, 2, 538–541. [Google Scholar] [CrossRef] [PubMed]
  44. Ortega, N.; Richter, C.; Glorius, F. N-Formylation of amines by methanol activation. Org. Lett. 2013, 15, 1776–1779. [Google Scholar] [CrossRef] [PubMed]
  45. Tumma, H.; Nagaraju, N.; Reddy, K.V. A facile method for the N-formylation of primary and secondary amines by liquid phase oxidation of methanol in the presence of hydrogen peroxide over basic copper hydroxyl salts. J. Mol. Catal. A Chem. 2009, 310, 121–129. [Google Scholar] [CrossRef]
  46. Han, Y.; Zhikang, W.; Zheyu, W.; Yongyan, Z.; Shi, R.; Qixin, Z.; Jingjing, W.; Sheng, H.; Yongge, W. N-formylation of amines using methanol as a potential formyl carrier by a reusable chromium catalyst. Commun. Chem. 2019, 2, 1–7. [Google Scholar] [CrossRef] [Green Version]
  47. Kaur, M.; Sharma, S.; Bedi, P.S. Silica supported Brönsted acids as catalyst in organic transformations: A comprehensive review. Cuihua Xuebao/Chinese J. Catal. 2015, 36, 520–549. [Google Scholar] [CrossRef]
  48. Pramanik, A.; Bhar, S. Silica-sulfuric acid and alumina-sulfuric acid: Versatile supported Brønsted acid catalysts. New J. Chem. 2021, 45, 16355–16388. [Google Scholar] [CrossRef]
  49. Habibi, D.; Rahmani, P.; Akbaripanah, Z. N-formylation of anilines with silica sulfuric acid under solvent free conditions. J. Org. Chem. 2013, 2013, 268654. [Google Scholar] [CrossRef]
  50. Hu, D.X.; Grice, P.; Ley, S.V. Rotamers or diastereomers? An overlooked NMR solution. J. Org. Chem. 2012, 77, 5198–5202. [Google Scholar] [CrossRef]
  51. Lanyon-Hogg, T.; Ritzefeld, M.; Masumoto, N.; Magee, A.I.; Rzepa, H.S.; Tate, E.W. Modulation of Amide Bond Rotamers in 5-Acyl-6,7-dihydrothieno [3,2-c]pyridines. J. Org. Chem. 2015, 80, 4370–4377. [Google Scholar] [CrossRef] [Green Version]
  52. Habibi, D.; Nasrollahzadeh, M.; Sahebekhtiari, H. Green synthesis of formamides using the Natrolite zeolite as a natural, efficient and recyclable catalyst. J. Mol. Catal. A Chem. 2013, 378, 148–155. [Google Scholar] [CrossRef]
  53. Ma’mani, L.; Sheykhan, M.; Heydari, A.; Faraji, M.; Yamini, Y. Sulfonic acid supported on hydroxyapatite-encapsulated-γ-Fe2O3 nanocrystallites as a magnetically Brønsted acid for N-formylation of amines. Appl. Catal. A Gen. 2010, 377, 64–69. [Google Scholar] [CrossRef]
  54. Bose, A.K.; Ganguly, S.N.; Manhas, M.S.; Guha, A.; Pombo-Villars, A. Microwave promoted energy-efficient N-formylation with aqueous formic acid. Tetrahedron Lett. 2006, 47, 4605–4607. [Google Scholar] [CrossRef]
  55. Lygin, A.V.; De Meijere, A. ortho-Lithiophenyl isocyanide: A versatile precursor for 3H-quinazolin-4-ones and 3H-quinazolin-4-thiones. Org. Lett. 2009, 11, 389–392. [Google Scholar] [CrossRef]
  56. Landquist, J.K. Synthetic antimalarials. Part XLVI. Some 4-[(dialkylaminoalkyl)amino] quinoline derivatives. J. Chem. Soc. 1951, 10, 1038–1048. [Google Scholar] [CrossRef]
  57. Kim, J.J.; Park, Y.D.; Cho, S.D.; Kim, H.K.; Chung, H.A.; Lee, S.G.; Falck, J.R.; Yoon, Y.J. Efficient N-arylation of pyridazin-3(2H)-ones. Tetrahedron Lett. 2004, 45, 8781–8784. [Google Scholar] [CrossRef]
  58. Trost, B.M. The Atom Economy—A Search for Synthetic Efficiency. Science 1991, 254, 1471–1477. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation depicting N-formamide as versatile synthetic reagent.
Figure 1. Schematic representation depicting N-formamide as versatile synthetic reagent.
Molecules 27 04213 g001
Figure 2. Immobilized sulfuric acid on silica gel.
Figure 2. Immobilized sulfuric acid on silica gel.
Molecules 27 04213 g002
Figure 3. N-formylation of amines with triethyl orthoformate.
Figure 3. N-formylation of amines with triethyl orthoformate.
Molecules 27 04213 g003
Figure 4. Proposed mechanism for N-formylation of amines with triethyl orthoformate.
Figure 4. Proposed mechanism for N-formylation of amines with triethyl orthoformate.
Molecules 27 04213 g004
Figure 5. Yield of isolated conformers 12A and 35A.
Figure 5. Yield of isolated conformers 12A and 35A.
Molecules 27 04213 g005
Table 1. Catalysts in combination with formylating agents employed for the formylation of various amines.
Table 1. Catalysts in combination with formylating agents employed for the formylation of various amines.
EntryCatalystFormylating AgentReaction ConditionTimeYield %Reference
1Sodium formateFormic acidSolvent free>8 h [31]
2Amberlite IR-120Formic acidMicrowave irradiation2 min90–97[23]
3Molecular iodine (I2)Formic acidSolvent free2 h60–99[32]
4Thiamine hydrochlorideFormic acidSolvent free 88–96[33]
5Fe2O3-Hap-SO3HFormic acidSolvent free15–60 min95–99[34]
6Sulfated tungstateFormic acidSolvent free10–45 min85–95[35]
7CDMT IIFormic acidMicrowave irradiation3–6 min64–94[36]
8Amidine and GuanidineMethyl formateSolvent free1–96 h65–98[37]
9TBD-based ionic liquidsFormic acidSolvent free10–35 min75–98[38]
10IndiumFormic acidSolvent free1.5–24 h70–98[28]
11ZnOFormic acidSolvent free10–720 min65–99[24]
12ZnCl2Formic acidSolvent free10–900 min60–98[39]
13TiO2-P25 or TiO2-SO42−Formic acidSolvent free30–45 min40–99[29]
14FSG-HF(N(SO2C8F11)2)4Formic acidSolvent free1–4 h60–88[40]
15IridiumParaformaldehydeReflux in H2O5–10 h41–91 [41]
16Silver and gold surfacesFormaldehydeSolvent free6 h75–97[42]
17Gold nanoparticles (Au/Al2O3 or Au/NiO)MethanolReflux in H2O4 h72–97[43]
18Ruthenium N-heterocyclic catalyst (Ru-NHC)MethanolReflux in toluene (125 °C)12–24 h27–99[44]
19Copper salt (CuCl2.H2O)MethanolSolvent free45–90 min63–80[45]
20Ionic liquid catalyzed formylationCOReflux in methanol (140 °C) 4 h42–99[18]
21Inorganic ligand-supported chromium (III) catalyst (NH4)3[CrMo6O18(OH)6]MethanolReflux in H2O2 (80 °C)12 h60–99[46]
22LipaseEthyl formateReflux in THF at room temperature1–8 h29–99[14]
23No catalystTriethyl orthoformate in waterUltrasound irradiation3 h35–88[1]
24Catalyst freeAmmonium formateSolvent free5 min–24 h43–98[13]
Table 2. Silica-supported Brønsted acids as catalyst for the formylation of various amines.
Table 2. Silica-supported Brønsted acids as catalyst for the formylation of various amines.
EntryCatalystFormylation AgentReaction ConditionTimeYield %Reference
1HClO4–SiO2Formic acidSolvent free15–90 min70–96[25]
2Fe3O4@SiO2–APTES-TFA1,3-dicarbonyl compoundSolvent freen/a68–98[34]
3H2SO4–SiO2Formic acidSolvent free4–46 min65–99[20]
4H2SO4–SiO2N,N-dimethyl amideSolvent free6–12 h75–95[25]
n/a: not applicable.
Table 3. Optimization of reaction parameters for N-formylation of amines with triethyl orthoformate (TEOF).
Table 3. Optimization of reaction parameters for N-formylation of amines with triethyl orthoformate (TEOF).
EntryReaction ConditionTimeYield
1Aniline (1 mmol)/TEOF (1 mmol), SIS (0.2 g)10 min44%
2Aniline (1 mmol)/TEOF (2 mmol), SIS (0.2 g)6 min66%
3Aniline (1 mmol)/TEOF (3 mmol), SIS (0.2 g)4 min96%
4Aniline (1 mmol)/TEOF (4 mmol), SIS (0.2 g)4 min90%
Table 4. N-formylation of aniline under different catalytic conditions.
Table 4. N-formylation of aniline under different catalytic conditions.
EntryCatalytic ConditionTime Yield
1Aniline (1 mmol)/TEOF (3 mmol) without catalyst at 65 °C3 htraces
2Aniline (1 mmol)/TEOF (3 mmol), SIS (0.1 g), 65 °C5 min78%
3Aniline (1 mmol)/TEOF (3 mmol), SIS (0.2 g), 65 °C4 min96%
4Aniline (1 mmol)/TEOF (3 mmol), SIS (0.3 g), 65 °C4 min88%
5Aniline (1 mmol)/TEOF (3 mmol), SIS (0.4 g), 65 °C6 min71%
6Aniline (1 mmol)/TEOF (3 mmol), SIS (0.5 g), 65 °C6 min64%
Table 5. N-formylation of amines using triethyl orthoformate in the presence of immobilized sulfuric acid on silica gel.
Table 5. N-formylation of amines using triethyl orthoformate in the presence of immobilized sulfuric acid on silica gel.
EntryAminesTime (Min)ProductYield (%)
1 Molecules 27 04213 i0014 Molecules 27 04213 i00296
2 Molecules 27 04213 i0034 Molecules 27 04213 i00481
3 Molecules 27 04213 i0054 Molecules 27 04213 i00678
4 Molecules 27 04213 i0079 Molecules 27 04213 i00895
5 Molecules 27 04213 i0094 Molecules 27 04213 i01090
6 Molecules 27 04213 i0114 Molecules 27 04213 i01297
7 Molecules 27 04213 i01310 Molecules 27 04213 i01483
8 Molecules 27 04213 i01510 Molecules 27 04213 i01697
9 Molecules 27 04213 i01710 Molecules 27 04213 i01890
10 Molecules 27 04213 i01910 Molecules 27 04213 i02096
11 Molecules 27 04213 i02115 Molecules 27 04213 i02290
12 Molecules 27 04213 i02313 Molecules 27 04213 i02475
13 Molecules 27 04213 i02513 Molecules 27 04213 i02681
14 Molecules 27 04213 i0275 Molecules 27 04213 i02886
15 Molecules 27 04213 i0295 Molecules 27 04213 i03094
16 Molecules 27 04213 i03120 Molecules 27 04213 i03275
17 Molecules 27 04213 i03312 Molecules 27 04213 i03473
18 Molecules 27 04213 i03520 Molecules 27 04213 i03685
19 Molecules 27 04213 i0376 Molecules 27 04213 i03897
20 Molecules 27 04213 i0396 Molecules 27 04213 i04078
21 Molecules 27 04213 i0415 Molecules 27 04213 i04294
22 Molecules 27 04213 i0436 Molecules 27 04213 i04478
23 Molecules 27 04213 i0456 Molecules 27 04213 i04684
24 Molecules 27 04213 i0475 Molecules 27 04213 i04881
25 Molecules 27 04213 i04910 Molecules 27 04213 i05056
26 Molecules 27 04213 i05110 Molecules 27 04213 i05281
27 Molecules 27 04213 i05312 Molecules 27 04213 i05482
28 Molecules 27 04213 i05515 Molecules 27 04213 i05685
29 Molecules 27 04213 i05715 Molecules 27 04213 i05896
30 Molecules 27 04213 i0598 Molecules 27 04213 i06093
31 Molecules 27 04213 i0616 Molecules 27 04213 i06294
32 Molecules 27 04213 i06320 Molecules 27 04213 i06496
33 Molecules 27 04213 i06518 Molecules 27 04213 i06695
34 Molecules 27 04213 i0675 Molecules 27 04213 i06886
35 Molecules 27 04213 i06912 Molecules 27 04213 i07093
36 Molecules 27 04213 i07112 Molecules 27 04213 i07298
37 Molecules 27 04213 i07315 Molecules 27 04213 i07480
38 Molecules 27 04213 i07520 Molecules 27 04213 i07691
39 Molecules 27 04213 i07724 Molecules 27 04213 i07893
40 Molecules 27 04213 i07915 Molecules 27 04213 i08095
41 Molecules 27 04213 i08113 Molecules 27 04213 i08292
42 Molecules 27 04213 i08325 Molecules 27 04213 i08477
43 Molecules 27 04213 i08530 Molecules 27 04213 i08667
44 Molecules 27 04213 i08754 Molecules 27 04213 i08876
45 Molecules 27 04213 i08945 Molecules 27 04213 i09079
46 Molecules 27 04213 i09145 Molecules 27 04213 i09271
47 Molecules 27 04213 i09360 Molecules 27 04213 i09494
48 Molecules 27 04213 i09550 Molecules 27 04213 i09694
49 Molecules 27 04213 i09740 Molecules 27 04213 i09887
50 Molecules 27 04213 i09940 Molecules 27 04213 i10078
51 Molecules 27 04213 i10150 Molecules 27 04213 i10273
52 Molecules 27 04213 i10340 Molecules 27 04213 i10485
53 Molecules 27 04213 i10540 Molecules 27 04213 i10675
54 Molecules 27 04213 i10760 Molecules 27 04213 i10885
55 Molecules 27 04213 i10935 Molecules 27 04213 i11093
56 Molecules 27 04213 i11160 Molecules 27 04213 i11293
Table 6. Efficiency of the recycled SIS in the N-formylation of aniline.
Table 6. Efficiency of the recycled SIS in the N-formylation of aniline.
EntryTurnYield %
1196
2289
3383
Table 7. Comparison of efficiency of various conditions in the N-formylation of aniline.
Table 7. Comparison of efficiency of various conditions in the N-formylation of aniline.
EntryConditionsTimeYieldReferences
1Triethyl orthoformate in H2O under ultrasound irradiation.3 h88%[1]
2Solid-supported formate, DMSO, 70–80 °C4 h60%[15]
3SSA, HCOOH, 50–60 °C, solvent-free7 min99%[49]
4SA on activated charcoal, ethylformate, 54 °C4 min95%[21]
5Triethyl orthoformate in H2O under neutral condition.
Microwave irradiation, 90 °C
2 h87%[2]
6SIS, triethyl orthoformate, 60–65 °C, solvent-free3 min96%Present protocol
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Salami, S.A.; Siwe-Noundou, X.; Krause, R.W.M. Catalytic Performance of Immobilized Sulfuric Acid on Silica Gel for N-Formylation of Amines with Triethyl Orthoformate. Molecules 2022, 27, 4213. https://doi.org/10.3390/molecules27134213

AMA Style

Salami SA, Siwe-Noundou X, Krause RWM. Catalytic Performance of Immobilized Sulfuric Acid on Silica Gel for N-Formylation of Amines with Triethyl Orthoformate. Molecules. 2022; 27(13):4213. https://doi.org/10.3390/molecules27134213

Chicago/Turabian Style

Salami, Sodeeq Aderotimi, Xavier Siwe-Noundou, and Rui W. M. Krause. 2022. "Catalytic Performance of Immobilized Sulfuric Acid on Silica Gel for N-Formylation of Amines with Triethyl Orthoformate" Molecules 27, no. 13: 4213. https://doi.org/10.3390/molecules27134213

Article Metrics

Back to TopTop