Vous êtes sur la page 1sur 4

View Online / Journal Homepage / Table of Contents for this issue

PAPER

www.rsc.org/greenchem | Green Chemistry

A Brnsted acidic ionic liquid as an efficient and reusable catalyst system for esterification{
Haibo Zhang, Fei Xu, Xiaohai Zhou,* Gaoyong Zhang and Cunxin Wang
Received 11th April 2007, Accepted 25th June 2007 First published as an Advance Article on the web 11th July 2007 DOI: 10.1039/b705480g Under mild conditions and without any additional organic solvent, esterification of alcohols by carboxylic acids could be carried out in a halogen-free Brnsted acidic ionic liquid, N-methyl-2pyrrolidonium methyl sulfonate ([NMP]+CH3SO32). Good conversion rates and high selectivities were obtained, especially for esterification of ethanol by acetic acid, and the liquid esters formed a separate phase that was decanted. The ionic liquid could be reused after removal of water under vacuum.

Downloaded by Fundacao Universidade do Rio Grande on 29 February 2012 Published on 11 July 2007 on http://pubs.rsc.org | doi:10.1039/B705480G

Introduction
Esterification is one of the most fundamental reactions in organic synthetic chemistry extensively employed for the protection and further manipulation of the carboxylic acid functional group.1,2 Esterification processes are widespread in the industrial synthesis of a variety of end-products, such as fragrances, monomers, solvents, and plasticizers etc., many of which are classed as high production volume (HPV) chemicals. In view of their importance, esterification protocols should occupy a prominent place in the desire to advance benign and sustainable chemical technologies into industrial process development. However, the esterification reaction is an equilibrium reaction and generally requires removal of water and/or use of excess amount of the reactants for satisfactory conversion rate.3 For the preparation of esters, the major drawbacks of these common methods are the final neutralization of the homogeneous acid catalyst, or separation of the metal catalyst supported on a solid support and, moreover, the removal of adsorbed products from the catalyst is quite difficult and requires a large excess of volatile organic solvents.46 These issues make this process environmentally questionable. The use of room temperature ionic liquids (ILs) as solvent and catalyst for chemical reactions offers many advantages from an environmental perspective.712 These solvents are nonflammable, thermally stable, exhibit negligible vapor pressure (non-volatile), and offer the potential for recyclability. For these reasons, the replacement of current esterification protocols with a more environmentally benign process involving the use of ionic liquids appeared to be an area worthy of investigation. Deng et al. firstly reported esterification of carboxylic acids with alcohols in acidic chloroaluminate ionic liquids.13 Since then, many Brnsted acidic ionic liquids, including SO3H-functionalized ionic liquids,1418 ionic liquids
College of Chemistry, Wuhan University, Wuhan, Hubei, P. R. China. E-mail: colloid@whu.edu.cn; Fax: +86 27 87218534; Tel: +86 27 87218534 { Electronic supplementary information (ESI) available: The NMR and ESI-MS spectra of [NMP]+CH3SO32 and the GC-MS spectra of esters. See DOI: 10.1039/b705480g

with acidic counteranion and protonated N-alkylimidazolium ionic liquids have been used in esterification reactions.19,20 Although elegant work has been done in this area, there still exist some drawbacks in the above catalytic systems. For example, the Brnsted IL [Hmim]+BF42 shows a satisfactory conversion rate and selectivity for esterification,20 but it needs a large amount of IL (e.g., 2 mL alcohol per 2 mL ionic liquid) and this ionic liquid contains halogen atom, which may cause serious concerns under certain conditions. Furthermore, the reaction temperature is high (110 uC). Therefore, further investigation is necessary to elucidate the merits of using Brnsted acidic ILs for esterification. In this paper, the halogen-free Brnsted acidic IL N-methyl2-pyrrolidonium methyl sulfonate ([NMP]+CH3SO32) was firstly synthesized (Scheme 1), and then used for esterification of alcohols by carboxylic acids (Scheme 2). [NMP]+CH3SO32 was demonstrated to be a very efficient and reusable catalyst systems for esterification. The IL could be a potential substitute for the prevalent esterification catalytic systems for the following reasons: (1) [NMP]+CH3SO32 shows superior catalytic activity than the systems reported previously, (2) synthesis and purification processes are easy and product reproducibility is good. (3) The costs of N-methyl-2-pyrrolidone as a source of cation are more economical compared to those of 1-methylpyrrolidine and 1-methylimidazole. Therefore, the use of [NMP]+CH3SO32 will reduce the production costs.

Scheme 1

Scheme 2

1208 | Green Chem., 2007, 9, 12081211

This journal is The Royal Society of Chemistry 2007

View Online

Experimental
General remarks All commercial chemicals were used as received. The reactions were carried out in a round-bottom flask with a stirrer. The reacting acids and alcohols (indicated in Table 1) were added to the IL [NMP]+CH3SO32. The esterification reactions proceed for a period of time ranging from 2 to 10 h with vigorous stirring at room temperature. The concentration of reactant and product was directly measured by a Finnigan Trace DSQ GC-MS equipped with a HP 5MS column (30 m long, 0.25 mm i.d., 0.25 mm film thickness) using the area of each chromatograph peak. After the reaction, the ester and ionic liquid [NMP] +CH3SO32 were separated conveniently by decanting, and the IL was reused after removal of water under vacuum (0.01 Torr) at 130 uC for 1 h. The water content of the IL was determined by coulometric KarlFischer titration using a WA-2 Coulometer, and duplicate measurements were performed on each sample with results agreeing to within 5%. Preparation of ionic liquid N-methyl-2-pyrrolidonium methyl sulfate ([NMP]+CH3SO32) The preparation of the Brnsted acidic IL was similar to those used in the previous literature21,22 and the detailed preparations are as follows. Benzene (30 mL) was added to a 50 mL flask containing 9.9 g of N-methyl-2-pyrrolidone (0.1 mol) and stirred. Then, 9.6 g of methane sulfonic acid (0.1 mol) was dripped slowly into the flask within ca. 30 min in an ice bath. The reaction lasted for another 4 h at room temperature. Benzene was removed under reduced pressure and further dried at 90 uC under 15 mmHg for 1 h. 1H-NMR (DMSO-d6, d/ppm relative to TMS): 15.342 (s, 1H), 2.8512.802 (t, 2H, J = 14.7 Hz), 2.524 (s, 3H), 2.226 2.184 (d, 3H, J = 12.6 Hz), 1.5961.571 (d, 4H, J = 7.5 Hz). ESI-MS: m/z = 198 [M + 2]+.

Typical esterification procedure 1-Butanol (7.4 g, 0.1 mol), equivalent acetic acid (6 g, 0.1 mol) and [NMP]+CH3SO32 (5 g, 0.025 mol) were added in a 50 mL flask with a stirrer. The reaction mixture was stirred for 4 h at room temperature. Reaction progress was monitored by GCMS. After reaction, the mixture was decanted, the ester was isolated, and the IL was reused after removal of water under vacuum (0.01 Torr) at 130 uC for 1 h.

Results and discussion


The fresh IL is a viscous colorless liquid at room temperature. However, it will slowly solidify at room temperature after a long time (e.g., several days). This phenomenon is consistent with the behavior of other ILs.23 The IL is miscible with water and readily soluble in methanol and other alcohols. Furthermore, it is partially immiscible with esters, alkanes and aromatic hydrocarbons. Because of this property, the IL can be separated conveniently from these solvents. Our initial efforts were directed toward investigating the formation of an important commodity ester,24 butyl acetate, from butanol and acetic acid using [NMP]+CH3SO32 as the solvent/catalyst in a batch-type process. The results of the experiments are summarized in Table 1. It was found that the IL had a very high activity for esterification. Excellent conversion rate and perfect selectivity were obtained in all cases at room temperature. Because of the high degree of miscibility of the IL with water and partial immiscibility with the produced ester, the esterification reaction equilibrium shifted to the product side even without simultaneous removal of the produced water, even though esterification is a reversible reaction. The effect of reaction time on conversion was investigated first. The results for the esterification of acetic acid with butanol (Table 1, entries 13) demonstrate that chemical equilibrium can be reached in 4 h at room temperature (just about 25 uC) without any other optimization. Increasing the reaction temperature (for example to 100 uC) and prolonging the reaction time did not offer significant advantages (Table 1, entry 4). Moreover, the ether (1.6%) and methane sulfonate (1.8%) was found as byproducts. The best reaction temperature was ca. 30 uC and the best reaction time was ca. 4 h, although the conversion rate of esterification would slightly increase with prolonged reaction time (Table 1, entries 13). Increasing the molar ratio of IL/substrate to 0.5 resulted in no significant increase in either conversion rate or selectivity (Table 1, entry 6). This indicated that the IL was an effective catalyst for esterification. Increasing the molar ratio of acid/ butanol to 1.2 resulted in a significant increase in conversion rate (Table 1, entries 7 and 8) and no butanol remained. Furthermore, for the purpose of comparison, the same esterification reaction was also carried out with equivalent of methane sulfonic acid as catalyst (Table 1, entries 9 and 10). It could be seen that methane sulfonic acid shows some catalytic activity for esterification. Prolonging the reaction time (Table 1, entry 10) and increasing the temperature (Table 1, entry 11) did not offer significant advantages. Furthermore, the ether (2.3%) and methane sulfonate (2.8%) were found as byproducts.
Green Chem., 2007, 9, 12081211 | 1209

Downloaded by Fundacao Universidade do Rio Grande on 29 February 2012 Published on 11 July 2007 on http://pubs.rsc.org | doi:10.1039/B705480G

Table 1 Results of esterification butanol by acetic acid under different conditionsa Acid : butanol : IL (mol : mol : mol) 4:4:1 4:4:1 4:4:1 4:4:1 10 : 10 : 1 4:4:2 4.8 : 4 : 1 12 : 10 : 1 Conversion (%) 90 92 95 97 95 96 99 99 75 75.6 68 Selectivity (%) 100 100 100 98.5 100 100 100 100 100 100 93 Temp./ uC r.t. r.t. r.t. 100 r.t. r.t. r.t. r.t. r.t. r.t. 100

Entry 1 2 3 4 5 6b 7b 8b 9c 10c 11c

Time/h 1 2 4 4 10 4 4 10 4 10 4

a Selectivity and conversion rate are based on GC-MS, no unreacted substrate and product in the IL phase were detected by GC-MS. The water content of [NMP]+CH3SO32 is 4.3 wt%. b Selectivity and conversion rate (butanol) are based on GC-MS. c Concentrated methane sulfonic acid 4.8 g (0.05 mol), ethanol 0.2 mol, acetic acid 0.2 mol.

This journal is The Royal Society of Chemistry 2007

View Online

From Table 1, it can be seen that the catalytic performance of the IL could be much better than that of the methane sulfonic acid under the same or more mild reaction conditions. This demonstrated that it was the Brnsted acidic IL rather than impurities (acid) that catalyzed this reaction. The product, butyl acetate, could be separated conveniently by decanting from the ionic liquid phase and purified by distillation for any leaching of the IL into the product phase. In order to investigate the scope and limitation of the [NMP]+CH3SO32 as a catalyst for esterification, other acids and alcohols as the substrates were also tested, and the results are summarized in Table 2. It was found that the IL [NMP]+CH3SO32 was of very high activity for esterification. Excellent conversion rates and perfect selectivity were obtained in all cases. Equivalent ethanols and acids were smoothly transformed to esters in high yield. No by-products, such as olefins or ethers, were detected. And all of the esters produced could be easily separated due to their immiscibility with the IL. It is noteworthy that removal of water or using an excess of reactant is not necessary. No volatile organic solvents are required during the work-up and recycling of the IL [NMP]+CH3SO32. The results in Table 2 showed that the esterification of aliphatic acids with alcohols was very satisfactory. The length of alkyl chains of alcohols did not affect the conversion and the selectivity (Table 2, entries 17). In the reaction of acetic acid and 1,4-butanediol, no monoesterification products were detected when the molar ratio of acid to alcohols was 2 : 1 (Table 2, entry 11). And in the reaction of oxalic acid, no monoesterification products were detected when the molar ratio of acid to alcohols was 1 : 2 (Table 2, entry 12). The hydroxy group of lactic acid did not take part in the esterification, with only 1-butyl lactate being obtained (Table 2, entry 13). Esterification of aromatic acids or aromatic alcohols was more difficult compared with that of aliphatic acids and longer reaction times were used, which gave
Table 2 Entry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

lower yields compared with those of aliphatic acids (Table 2, entries 8, 14 and 15). The esterification of long alkyl chain acids was not very easy and the required reaction time was longer than that of the other reaction (Table 2, entries 16, 17, 18 and 19). However, satisfactory results were obtained. Olefinic bonds of acids were not changed under these reaction conditions (Table 2, entries 14 and 19). Once the reactions were complete, more attention was paid to isolating the products and recycling the catalyst. The water produced in the reactions did not need to be removed because the IL [NMP]+CH3SO32 was miscible with water while the esters were immiscible with the IL. Hence, the esterification proceeded smoothly to completion. Liquid esters could be separated conveniently by decanting. Some esters (Table 2, entry 18) have to be isolated above their melting point since they solidify at room temperature. In order to investigate the possibility of recycling of the Brnsted acidic IL [NMP]+CH3SO32, a recycling experiment was conducted. For each cycling reaction, butanol (14.8 g, 0.2 mol), acetic acid (12 g, 0.2 mol) and [NMP]+CH3SO32 (5 g, 0.05 mol) were added in a round-bottom flask successively and reacted at room temperature for 4 h. After reaction, the product could be conveniently decanted out from the IL. After being treated under vacuum (0.01 Torr) at 130 uC for 1 h, the IL was assessed by 1H NMR spectroscopy and no traces of acetic and ethanol were detected. The water content of [NMP]+CH3SO32 is 4.8 wt%. The reactor containing the used IL was charged with acetic and butanol again after drying under vacuum. This procedure was repeated for six cycles. The conversion rate and selectivity are listed in Table 3. It could be seen that the catalytic activity was slightly decreased after [NMP]+CH3SO32 was used for five times. This indicated that the IL as catalyst for the esterification was recyclable. The slight decrease of conversion rate might be ascribed to the slight deactivation of [NMP]+CH3SO32.

Downloaded by Fundacao Universidade do Rio Grande on 29 February 2012 Published on 11 July 2007 on http://pubs.rsc.org | doi:10.1039/B705480G

Results of esterification for different acids and alcohols in [NMP]+CH3SO32 a Acid Acetic acid Acetic acid Acetic acid Acetic acid Acetic acid Acetic acid Acetic acid Acetic acid Propionic acid Propionic acid Acetic acid Oxalic acid Lactic acid Cinnamic acid Benzoic acid Lauryl acid Lauryl acid Stearic acid Oleic acid Alcohol Propanol Isopropanol 1-Butanol Isoamylol 1-Pentanol 1-Hexanol 1-Heptanol Benzyl alcohol Ethanol 1-Butanol 1,4-Butanediol 1-Butanol 1-Butanol Ethanol Ethanol Methanol Ethanol Methanol Methanol Conversionb (%) 95 94 95 92 95 96 94 85 94 93 95 96 94 84 86 96 95 92 93 Selectivity to ester (%) 100 100 100 100 100 100 100 100 100 100 100c 100c 100 100 100 100 100 100 100 Temp./uC r.t. r.t. r.t. r.t. r.t. r.t. r.t. 60 r.t. r.t. 60 60 60 60 60 r.t. r.t. 40 r.t. Time/h 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 10 10 10 10

a Selectivity and conversion rate are based on GC-MS, no unreacted substrate and product in IL phase were detected by GC-MS. b The conversion rate of acid or alcohol (mole ratio of acid to alcohol 1 : 1); the molar ratio of IL to acid or alcohol 1 : 4. c The esters are all double esterification products. No monoesterification product was detected by GC-MS; the mole ratio of acid to alcohol = 2 : 1 in entry 11, and ratio of acid to alcohol = 1 : 2 in entry 12

1210 | Green Chem., 2007, 9, 12081211

This journal is The Royal Society of Chemistry 2007

View Online

Table 3 Recycling of [NMP]+CH3SO32 in the synthesis of butyl acetate Run 1 2 3 4 5 6 Conversion (%) 95 94 95 96 95 93 Selectivity to ester (%) 100 100 100 100 100 100 Time/h 4 4 4 4 4 4

References
1 M. B. Smith and J. March, Advanced Organic Chemistry, Wiley, New York, USA, 5th edn, 2001. 2 R. C. Larock, Comprehensive Organic Transformations, WileyVCH, New York, USA, 2nd edn, 1999. 3 A. S. Y. Lee, H. C. Yang and F. Y. Su, Tetrahedron Lett., 2001, 42, 301. 4 P. Maki-Arfela, T. Salmi, M. Sundell, K. Ekman, R. Peltonen and J. Lehtonen, Appl. Catal., A, 1999, 184, 25. 5 M. Hino and K. Arata, Appl. Catal., 1985, 18, 401. 6 M. A. Schwegler, H. van Bekkum and N. A. de Munck, Appl. Catal., 1991, 74, 191. 7 P. Wasserscheid and T. Welton, Ionic Liquids in Synthesis, WileyVCH, Weinheim, Germany, 2003. 8 T. Welton, Chem. Rev., 1999, 99, 2071. 9 R. Sheldon, Chem. Commun., 2001, 2399. 10 C. M. Gordon, Appl. Catal., A, 2001, 222, 101. 11 D. Zhao, M. Wu, Y. Kou and E. Min, Catal. Today, 2002, 74, 157. 12 H. Zhao and S. V. Malhotra, Aldrichimica Acta, 2002, 35, 75. 13 Y. Deng, F. Shi, J. Peng and K. Qiao, J. Mol. Catal. A: Chem., 2001, 165, 33. 14 A. C. Cole, J. L. Jensen, I. Ntai, K. L. T. Tran, K. J. Weave, D. C. Forbes and J. H. Davis, J. Am. Chem. Soc., 2002, 124, 5962. 15 D. C. Forbes and K. J. Weaver, J. Mol. Catal. A: Chem., 2004, 214, 12. 16 J. H. Davis, Jr., Chem. Lett., 2004, 33, 1072. 17 K. Qiao and C. Yokoyama, Chem. Lett., 2004, 33, 472. 18 H. Xing, T. Wang, Z. Zhou and Y. Dai, Ind. Eng. Chem. Res., 2005, 44, 4147. 19 J. Fraga-Dubreuil, K. Bourahla, M. Rahmouni, J. P. Bazureau and J. Hamelin, Catal. Commun., 2002, 3, 185. 20 H. Zhu, F. Yang, J. Tang and M. He, Green Chem., 2003, 5, 38. 21 Z. Du, Z. Li, S. Guo, J. Zhang, L. Zhu and Y. Deng, J. Phys. Chem. B, 2005, 109, 19542. 22 X. Zhou, H. Zhang and G. Zhang, CN Pat., CN 200610124650.2, 2006. 23 C. P. Fredlake, J. M. Crosthwaitte, D. G. Hert, S. N. V. K. Aki and J. F. Brennecke, J. Chem. Eng. Data, 2004, 49, 954. 24 J. Otera, Angew. Chem., Int. Ed., 2001, 40, 2044.

Downloaded by Fundacao Universidade do Rio Grande on 29 February 2012 Published on 11 July 2007 on http://pubs.rsc.org | doi:10.1039/B705480G

Although a detailed reaction mechanism of esterification in the Brnsted acidic IL [NMP]+CH3SO32 is not clear at this stage, the preliminary experimental results show that esterification in the Brnsted acidic IL, as a novel environmental benign catalyst, is not only possible but also quite satisfactory. In summary, the esterification of carboxylic acid with alcohol, using the IL [NMP]+CH3SO32 as catalyst and solvent, has several advantages: (1) the IL [NMP]+CH3SO32, as a strong Brnsted acid, shows superior catalytic activity than the systems reported previously. (2) The esterification procedure can be achieved at room temperature. (3) The preparation of [NMP]+CH3SO32 was very simple and the cost is low. (4) The IL contains no halogen atoms such as F and Cl. (5) [NMP]+CH3SO32 could be directly reused after removal of water since no byproducts were produced in the reaction. (6) The esters produced can be isolated conveniently in high yields and purity.

Acknowledgements
The authors are grateful to National Natural Science Foundation of China (Grant No. 20233010 and 20573079) for the financial support.

This journal is The Royal Society of Chemistry 2007

Green Chem., 2007, 9, 12081211 | 1211

Vous aimerez peut-être aussi