Copper-catalyzed N-formylation of amines with CO2 under ambient conditions

Suqi Zhang ab, Qingqing Mei b, Hangyu Liu b, Huizhen Liu *b, Zepeng Zhang *a and Buxing Han b
aSchool of Materials Science and Technology, China University of Geosciences, Beijing, 100083, P. R. China
bBeijing National Laboratory for Molecular Sciences, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China. E-mail: liuhz@iccas.ac.cn; unite508@163.com

Received 27th February 2016 , Accepted 23rd March 2016

First published on 24th March 2016


Abstract

We carried out work on N-formylation of amines with CO2 and PhSiH3 to produce formamides catalyzed by a copper complex. It was found that the Cu(OAc)2–bis(diphenylphosphino)ethane (dppe) catalytic system was very efficient for these kind of reactions at room temperature and 1 atm CO2 with only 0.1 mol% catalyst loading.


Carbon dioxide (CO2) chemistry is attracting increasing attention in both academia and industry because CO2 is a cheap, non-toxic and abundant C1 building block.1,2 Various approaches have been explored for the conversion of CO2, and the catalytic reduction of CO2 to value-added compounds is one of the most attractive routes. Various commodity chemicals such as formic acid derivatives,3 methane,4 methanol5 and formaldehyde6 can be produced.

Formamides are a class of chemicals with wide applications in industry as solvents and raw materials for synthesis of other chemicals. An interesting route for the synthesis of formamides is the N-formylation of amines with CO2 in the presence of reducing agent. Hydrogen gas is the cleanest and most atom-economical reductant, but harsh reaction conditions, such as high reaction pressure and temperature, often prevents its broad application.7 Moreover, it is known that the activity of aromatic amines is poor when using H2 as the reducing agent. Hydrosilanes possess a reduction potential similar to H2 and a Si–H bond that is kinetically more reactive because of its polarity and lower bond dissociation energy. In addition, they circumvent the air and moisture sensitivity of aluminium and boron hydrides and so hydrosilanes are a kind of mild and easily handling reducing agent. Various organocatalysts8 and organometallic complexes9 have been utilized to catalyze the N-formylation of amines with CO2 using hydrosilanes as the reductant. This kind of transformation was first reported in 2012 using organic base 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) and 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD) as the catalyst at a reaction temperature of 100 °C.2b Other organic catalysts such as N-heterocyclic carbenes (NHCs)8a and ionic liquid 1-butyl-3-methylimidazolium chloride ([BMIm]Cl)8b were also used to catalyze this reaction. However, N-heterocyclic carbenes (NHCs) is sensitivity to moisture and the catalyst loading was high (5 mol%). The [BMIm]Cl loading was 100% and high CO2 pressure (1 MPa) was required for this transformation. 1,3,2-Diazaphospholene could catalyze the N-formylation reaction at room temperature and 1 atm CO2, but the catalyst loading was high (5 mol%).8c Alkyl bridged chelating bis(NHC) rhodium complexes (NHC = N-heterocyclic carbene) was highly effective for the reaction at a low catalyst loading and ambient temperature,9a but at high pressure. In addition, the metal rhodium is expensive. Fe(acac)2 with PP3 catalyst system could catalyze the N-formylation reaction with 5 mol% catalyst loading at room temperature.9b Copper with diphosphine ligand 1,2-bis(diisopropylphosphino)benzene catalyst has been widely used to catalyze the reduction of CO2.10 Copper with 1,2-bis(diisopropylphosphino)benzene catalyzed the N-formylation using polymethylhydrosiloxane (PMHS) as the reducing agent,9a and the TON is up to 11[thin space (1/6-em)]700 at 80 °C. However, the value of TON is lower at room temperature.11 It was well known that ligand is very important for catalytic reaction. Considering copper is cheaper, more efficient copper complex catalytic system for N-formylation of amines with CO2 should be explored. Up to now, the formylation reaction at ambient temperature and atmospheric pressure with a low catalyst loading has not been achieved due mainly to the inherent kinetic stability of CO2.

Herein, we conducted the first work to use Cu-based catalysts in the N-formylation of amines with CO2 and PhSiH3. It was found that the N-formylation reactions could be catalyzed efficiently by 1,2-bis(diphenylphosphino)ethane (dppe)–Cu(II) complex at room temperature and 1 atm CO2 with only 0.1 mol% catalyst loading for the conversion of a wide range of aliphatic and aromatic amines, and the yields of the desired products were very high.

We check the effect of different ligand–Cu(II) catalysts in the N-formylation of amines with CO2 and PhSiH3. It was found that the bite angles of the ligands can affect significantly the yields of the desired product. The angle bite β° is bigger, the yield of the desired product is lower. The N-formylation reactions could be catalyzed efficiently by (dppe)–Cu(II) complex at room temperature and 1 atm CO2 with only 0.1 mol% catalyst loading for the conversion of a wide range of aliphatic and aromatic amines, and the yields of the desired products were very high. And the TON can be reached to about 14[thin space (1/6-em)]500 at room temperature with 1 atm CO2 for 4 h and the TOF is about 3625 h−1.

Firstly, the effect of different ligand shown in Fig. 1 using Cu(OAc)2 as the metal precursor was checked. The results were shown in Table 1. We found that mono-phosphine ligands (Cy3P, Ph3P, Tppi) are inactive for the N-formylation reaction. Bis-phosphine ligands are all active, and the yield of the desired product is related with the angle bite of the ligands. When the angle bite β° is bigger, the yield of the desired product is lower. The angle bite β° of dppm is 84°, while the yield of the desired is also very low because the dppm more easily forms dinuclear species complex. The angle bite of dppe and dppb is similar, while the yield of the desired of the product is higher for dppe, maybe because the ligand dppe is more flexible. It is obviously that dppe is the best ligand among the ligands checked (Table 1).


image file: c6ra05199e-f1.tif
Fig. 1 Phosphine ligands employed for N-formylation reaction.
Table 1 The effect of bite angle of the ligandsa

image file: c6ra05199e-u1.tif

Ligand Bite angle, β° Yield of formanilide (%)
a Reaction conditions: aniline (1 mmol), Si–H (6 mmol), Cu(OAc)2 (1 μmol), ligand (1.2 μmol); N.R. = No reaction.
Cy3P N.R.
Ph3P N.R.
TPPi N.R.
Xantphos 102 (ref. 12) 11
dppb 87 (ref. 13) 23
dppm 84 (ref. 14) 59
dppe 89 (ref. 12) 96
dppbt 98 (ref. 13) 83


The performances of different metal salts were evaluated for the N-formylation of aniline in the presence of dppe (Table 1). Among the salts studied, [Cu(OAc)2] was very active, while other salts were not active (Table 2, entries 1–12 and 15). In addition, the reaction did not take place when only [Cu(OAc)2] or dppe was used (Table 2, entries 13 and 14), indicating that both the Cu salt and ligand were necessary. Solvent also affected the reaction significantly. Toluene is the best solvent for the reaction and the yield of 2a could be as high as 96% (Table 2, entry 15). In THF and 1,4-dioxane the yields of 2a were 76% and 84%, respectively (Table 1, entries 16 and 17). The reaction did not take place in other solvents such as dichloromethane, hexane and CH3CN (Table 2, entries 18 and 19). Finally, the reductant agents Ph2SiH2 and PMHS did not afford any desired product at the reaction conditions (Table 2, entries 20 and 21).

Table 2 Optimization of catalytic system and reaction conditiona

image file: c6ra05199e-u2.tif

Entry Metal precursor Reducing agent Solvent Yield (%)
a Reaction conditions: aniline (1 mmol), Si–H (6 mmol), metal salt (1 μmol), ligand (1.2 μmol). b Without Cu(OAc)2·2H2O. c Without ligand.
1 AgCl phSiH3 Toluene <5%
2 AgOAc phSiH3 Toluene <5%
3 NiCl2·6H2O phSiH3 Toluene <5%
4 Ni(OAc)2 phSiH3 Toluene <5%
5 FeCl2 phSiH3 Toluene <5%
6 ZnCl2·6H2O phSiH3 Toluene <5%
7 Zn(OAc)2 phSiH3 Toluene <5%
8 CuCl2·2H2O phSiH3 Toluene <5%
9 CuBr2 phSiH3 Toluene <5%
10 CuBr phSiH3 Toluene <5%
11 CuI phSiH3 Toluene <5%
12 Cu(OTf)2 phSiH3 Toluene <5%
13b phSiH3 Toluene <5%
14c Cu(OAc)2·2H2O phSiH3 Toluene <5%
15 Cu(OAc)2·2H2O phSiH3 Toluene 96
16 Cu(OAc)2·2H2O phSiH3 THF 76
17 Cu(OAc)2·2H2O phSiH3 1,4-Dioxane 84
18 Cu(OAc)2·2H2O phSiH3 CH2Cl2 <5%
19 Cu(OAc)2·2H2O phSiH3 Hexane <5%
20 Cu(OAc)2·2H2O Ph2SiH2 Toluene <5%
21 Cu(OAc)2·2H2O PMHS Toluene <5%


Having in hand an efficient catalytic system for the hydroformylation of 1a, the scope of reactive amines were explored utilizing Cu(OAc)2·H2O and dppe (0.1 mol%) as a catalytic system with PhSiH3 as a reductant in toluene, and the results are listed in Table 3. It was demonstrated that different kinds of amines were formylated to the corresponding formamides in excellent yields at 25 °C and 1 atm CO2. The yields of the corresponding products were 96% and 97% respectively for aniline, 4-methyl aniline in 4 hours (Table 3, entries 1 and 2). When the substituent group was chloride, the reaction time need to be prolonged to 5 hours to get 95% yield of N-(4-chlorophenyl)formamide (Table 3, entry 3). Longer reaction time (6 h) was required for achieving high yield of N-mesitylformamide (Table 3, entry 4) because of steric hindrance. Primary aliphatic amine also showed high activity, while aliphatic amine with alkyl chains showed lower activity compared to benzyl amine and cyclohexanamine (Table 3, entries 5–7). All the primary amines checked only yielded monoformylated products (Table 3, entries 1–7). We did not detect any diformylated products using N-phenylformamide as the starting material. The yield of the N,N-diphenylformamide was 91% in 6 hours (Table 3, entry 11). N-Methylaniline, indoline and morpholine exhibited the highest activity, and the yields of the corresponding products could reach 99% in 2 hours (Table 3, entries 8–10). Longer reaction time was needed to get high yield of the product for dihexylamine and the yield of N,N-dihexylformamide could be increased to 96% after a longer reaction time (Table 3, entry 13). Secondary amines with alkyl chains showed lower activity compared to N-methylanilines and aliphatic primary amines (Table 3, entries 6, 8 and 13).

Table 3 N-Formylation of various amines with CO2a

image file: c6ra05199e-u3.tif

Entry Substrate Product Time (h) Yield (%)
a Reaction conditions: amine (1 mmol), PhSiH3 (2 mmol), Cu(OAc)2·2H2O (1 μmol), dppe (1.2 μmol).
1 image file: c6ra05199e-u4.tif image file: c6ra05199e-u5.tif 4 96
2 image file: c6ra05199e-u6.tif image file: c6ra05199e-u7.tif 4 97
3 image file: c6ra05199e-u8.tif image file: c6ra05199e-u9.tif 5 95
4 image file: c6ra05199e-u10.tif image file: c6ra05199e-u11.tif 8 92
5 image file: c6ra05199e-u12.tif image file: c6ra05199e-u13.tif 4 99
6 image file: c6ra05199e-u14.tif image file: c6ra05199e-u15.tif 5 95
7 image file: c6ra05199e-u16.tif image file: c6ra05199e-u17.tif 4 92
8 image file: c6ra05199e-u18.tif image file: c6ra05199e-u19.tif 2 99
9 image file: c6ra05199e-u20.tif image file: c6ra05199e-u21.tif 2 99
10 image file: c6ra05199e-u22.tif image file: c6ra05199e-u23.tif 2 99
11 image file: c6ra05199e-u24.tif image file: c6ra05199e-u25.tif 6 91
12 image file: c6ra05199e-u26.tif image file: c6ra05199e-u27.tif 4 99
13 image file: c6ra05199e-u28.tif image file: c6ra05199e-u29.tif 6 96


The possible reaction mechanism was shown in Scheme 1. The copper catalyst activated PhSiH3 to form B and catalyzed the insertion of CO2 into Si–H bond to form intermediate C, which further reacted with the amine, producing the final product. To get evidence for supporting this reaction mechanism, control experiment was performed. The reaction of PhSiH3 and CO2 was allowed to proceed for 1 h in the presence of the catalyst. The CO2 was removed and aniline was added and stirred for 4 h, and N-phenylformamide was also produced. This indicates that an intermediate was formed in the N-formylation reaction. To identify the intermediate, the mixture of catalyst, PhSiH3 and CO2 after reaction for 1 h without aniline was analyzed by 1H NMR and 13C NMR (Fig. 2 and 3). A signal appeared at δ = 163.0 ppm (Fig. 2), and a new one signal also appeared in the 1H NMR spectrum at δ = 8.18 ppm (Fig. 3). These two signals were ascribed to formoxysilane (Fig. 3), indicating that the reaction of phenylsilane with CO2 to form formoxysilane in the presence of the copper-based catalyst. After the addition of aniline, the formoxysilane intermediate transformed into the corresponding product. The mixture of catalyst, PhSiH3 and CO2 stirred 1 h, and then analyzed by GC-MS. Formoxysilane intermediate was detected (Fig. S1).


image file: c6ra05199e-s1.tif
Scheme 1 The possible reaction mechanism.

image file: c6ra05199e-f2.tif
Fig. 2 13C NMR spectra of catalyst, PhSiH3 and CO2 (DMSO-d6, 298 K).

image file: c6ra05199e-f3.tif
Fig. 3 1H NMR spectra of catalyst, PhSiH3 and CO2 (d6-DMSO, 298 K).

Conclusions

In conclusion, Cu(II)–dppe complex can efficiently catalyze the N-formylation reactions of various amines with CO2 and PhSiH3 at ambient condition with only 0.1 mol% catalyst loading. The yields of the corresponding desired products range from 91% to 99%, and no by-product is formed in the reactions. In a reaction, phenylsilane reacts with CO2 to form an intermediate formoxysilane, which reacts with the amine to yield the desired product. We believe that this route has great potential of application because the reactions proceed at ambient condition with very high efficiency, and the catalyst is cheap.

Acknowledgements

This work was supported by the Recruitment Program of Global Youth Experts of China, Chinese Academy of Sciences (KJCX2.YW.H30). National Natural Science Foundation of China (21533011, 21321063), and. The measurements of 1H and 13C nuclear magnetic resonance (NMR) and high resolution mass spectrometry (HRMS) were performed at the Centre for Physicochemical Analysis and Measurements in ICCAS.

Notes and references

  1. (a) C. Mandil, Tracking industrial energy efficiency and CO2 emissions, IEA Publications, Paris, France, 2007 Search PubMed; (b) K. M. K. Yu, I. Curcic, J. Gabriel and S. C. E. Tsang, ChemSusChem, 2008, 1, 893 CrossRef CAS PubMed; (c) M. Y. He, Y. H. Sun and B. X. Han, Angew. Chem. Int. Ed., 2013, 52, 9620 ( Angew. Chem. , 2013 , 125 , 9798 ) CrossRef CAS PubMed.
  2. (a) M. Aresta, Carbon Dioxide as Chemical Feedstock, ed. M. Aresta, Wiley-VCH, Weinheim, 2010, p. 1 CrossRef; (b) C. D. Gomes, O. Jacquet, C. Villiers, P. Thuery, M. Ephritikhine and T. Cantat, Angew. Chem. Int. Ed., 2012, 51, 187 ( Angew. Chem. , 2011 , 124 , 191 ) CrossRef PubMed; (c) K. Huang, C. L. Sun and Z. J. Shi, Chem. Soc. Rev., 2011, 40, 2435 RSC; (d) T. Sakakura, J. C. Choi and H. Yasuda, Chem. Rev., 2007, 107, 2365 CrossRef CAS PubMed; (e) P. G. Jessop, F. Joo and C. C. Tai, Coord. Chem. Rev., 2004, 248, 2425 CrossRef CAS; (f) W. Wang, S. P. Wang, X. B. Ma and J. L. Gong, Chem. Soc. Rev., 2011, 40, 3703 RSC.
  3. (a) X. D. Xu and J. A. Moulijn, Energy Fuels, 1996, 10, 305 CrossRef; (b) Z. F. Zhang, S. Q. Hu, J. L. Song, W. L. Li, G. Y. Yang and B. X. Han, ChemSusChem, 2009, 2, 234 CrossRef CAS PubMed; (c) T. C. Johnson, D. J. Morris and M. Wills, Chem. Soc. Rev., 2010, 39, 81 RSC; (d) Y. Himeda, Eur. J. Inorg. Chem., 2007, 7, 3927 CrossRef; (e) T. Matsuo and H. Kawaguchi, J. Am. Chem. Soc., 2006, 128, 12362 CrossRef CAS PubMed; (f) C. Y. Wu, Z. F. Zhang, Q. G. Zhu, H. L. Han, Y. Y. Yang and B. X. Han, Green Chem., 2015, 17, 1467 RSC; (g) K. Motokura, D. Kashiwame, A. Miyaji and T. Baba, Org. Lett., 2012, 14, 2642 CrossRef CAS; (h) Z. P. Lu, H. Hausmann, S. Becker and H. A. Wegner, J. Am. Chem. Soc., 2015, 137, 5332 CrossRef CAS PubMed.
  4. (a) A. Berkefeld, W. E. Piers and M. Parvez, J. Am. Chem. Soc., 2010, 132, 10660 CrossRef CAS PubMed; (b) S. Park, D. Bezier and M. Brookhart, J. Am. Chem. Soc., 2012, 134, 11404 CrossRef CAS PubMed; (c) R. J. Wehmschulte, M. Saleh and D. R. Powell, Organometallics, 2013, 32, 6812 CrossRef CAS.
  5. (a) S. N. Riduan, Y. G. Zhang and J. Y. Ying, Angew. Chem. Int. Ed., 2009, 48, 3322–3325 ( Angew. Chem. , 2009 , 121 , 3372 ) CrossRef CAS PubMed; (b) M. Courtemanche, M. Legare, L. Maron and F. G. Fontaine, J. Am. Chem. Soc., 2013, 135, 9326 CrossRef CAS PubMed; (c) N. Pasupuletya, H. Drissb, Y. A. Alhameda, A. A. Alzahrania, M. A. Daousa and L. Petrov, Appl. Catal., A, 2015, 504, 308 CrossRef; (d) F. Studt, I. Sharafutdinov, F. A. Pedersen, C. F. Elkjær, J. S. Hummelshøj, S. Dahl, I. Chorkendorff and J. K. Nørskov, Nat. Chem., 2014, 6, 320 CrossRef CAS PubMed.
  6. S. Bontemps, L. Vendier and S. Sabo-Etienne, J. Am. Chem. Soc., 2014, 136, 4419 CrossRef CAS PubMed.
  7. (a) X. J. Cui, Y. Zhang, Y. G. Deng and F. Shi, Chem. Commun., 2014, 50, 189 RSC; (b) L. Zhang, Z. B. Han, X. Y. Zhao, Z. Wang and K. L. Ding, Angew. Chem. Int. Ed., 2015, 54, 6186 ( Angew. Chem. , 2015 , 127 , 6284 ) CrossRef CAS; (c) L. Schmid, A. Canonica and A. Baiker, Appl. Catal., A, 2003, 255, 23 CrossRef CAS.
  8. (a) O. Jacquet, C. D. N. Gomes, M. Ephritikhine and T. Cantat, J. Am. Chem. Soc., 2012, 134, 2934 CrossRef CAS PubMed; (b) L. D. Hao, Y. F. Zhao, B. Yu, Z. Z. Yang, H. Y. Zhang, B. X. Han, X. Gao and Z. M. Liu, ACS Catal., 2015, 5, 4989 CrossRef CAS; (c) C. C. Chong and R. Kinjo, Angew. Chem. Int. Ed., 2015, 54, 12116–12120 ( Angew. Chem. , 2015 , 127 , 12284 ) CrossRef CAS PubMed.
  9. (a) X. Frogneux, O. Jacquet and T. Cantat, Catal.: Sci. Technol., 2014, 4, 1529 RSC; (b) T. V. Q. Nguyen, W. J. Yoo and S. Kobayashi, Angew. Chem. Int. Ed., 2015, 54, 9209 ( Angew. Chem. , 2015 , 127 , 9341 ) CrossRef CAS PubMed.
  10. K. Motokura, N. Takahashi, D. Kashiwame, S. Yamaguchi, A. Miyajiand and T. Baba, Catal.: Sci. Technol., 2013, 3, 2392 RSC.
  11. K. Motokura, N. Takahashi, A. Miyaji, Y. Sakamoto, S. Yamaguchi and T. Baba, Tetrahedron, 2014, 70, 6951 CrossRef CAS.
  12. M. Tromp, J. A. van Bokhoven, G. P. F. van Strijdonck, P. W. N. M. van Leeuwen, D. C. Koningsberger and D. E. Ramaker, J. Am. Chem. Soc., 2004, 127, 777 CrossRef PubMed.
  13. P. Comba, C. Katsichtis, B. Nuber and H. Pritzkow, Eur. J. Inorg. Chem., 1999, 777 CrossRef CAS.
  14. O. Moudam, A. Kaeser, B. Delavaux-Nicot, C. Duhayon, M. Holler, G. Accorsi, N. Armaroli, I. Séguy, J. Navarro, P. Destruel and J. F. Nierengarten, Chem. Commun., 2007, 3077 RSC.

Footnote

Electronic supplementary information (ESI) available: Experimental procedures, spectroscopic data, and supporting figures and tables. See DOI: 10.1039/c6ra05199e

This journal is © The Royal Society of Chemistry 2016