Propylene oxide as a dehydrating agent: potassium carbonate-catalyzed carboxylative cyclization of propylene glycol with CO2 in a polyethylene glycol/CO2 biphasic system

Zhen-Feng Diao a, Zhi-Hua Zhou a, Chun-Xiang Guo a, Bing Yu a and Liang-Nian He *ab
aState Key Laboratory and Institute of Elemento-Organic Chemistry, Nankai University, Tianjin 300071, China. E-mail: heln@nankai.edu.cn; Fax: +86-22-23503878; Tel: +86-22-23503878
bCollaborative Innovation Center of Chemical Science and Engineering, Tianjin 300071, China

Received 18th February 2016 , Accepted 22nd March 2016

First published on 24th March 2016


Abstract

The synthesis of propylene carbonate (PC) from 1,2-propylene glycol (PG) and CO2 was smoothly performed in a PEG800 (polyethylene glycol)/CO2 biphasic system with K2CO3 as a catalyst and propylene oxide (PO) as a dehydrating agent. In the reaction of PG with CO2, PO presumably removes the water produced, and simultaneously generates more PG, both of which shift the thermodynamic control process and thus accelerate the PC synthesis. The PC yield directly from PG and CO2 reached 78% under relatively mild reaction conditions (4 MPa, 120 °C, 10 h). Notably, no additional by-product was detected in this process, resulting in economic benefits and the ease of workup procedure.


Introduction

Ethylene carbonate (EC), propylene carbonate (PC) and dimethyl carbonate (DMC) are important industrial organic carbonates,1 which have been widely applied as polar aprotic solvents,2–4 electrolytes for lithium ion batteries, raw materials for polycarbonates/polyurethanes,5 green carbonylation reagents6 and fuel additives.7 Generally, PC and EC have also been used to prepare DMC through the transesterification with methanol.8

Currently, there are two conventional accesses to produce DMC: (a) phosgenation of methanol;9 (b) oxidative carbonylation of methanol with carbon monoxide (Scheme 1a and b).10 Corrosive and toxic reagents are commonly required in those procedures. As the most straightforward and challenging method for direct DMC synthesis, the dehydrating condensation of methanol with CO2 is highly attractive (Scheme 1c). Although DMC synthesis directly from CO2 and methanol catalyzed by transition metals has been developed, the process has not been yet applied for industrial application because of low productivity originating from catalyst deactivation and DMC hydrolysis caused by co-produced H2O. Therefore, considerable efforts have been devoted to overcoming such limitations.11–13 On the other hand, one of the most efficient and practicable strategies is the two-step process consisting of cycloaddition of epoxides with CO2, and subsequent transesterification of cyclic carbonate e.g. PC with methanol (Scheme 2a), which has widely been applied in industrial DMC production.14


image file: c6ra04422k-s1.tif
Scheme 1 DMC synthetic methodologies.

image file: c6ra04422k-s2.tif
Scheme 2 Two-step process for DMC production.

In this process, however, an equivalent of propylene glycol (PG) is inevitably released as co-product.15 Therefore, if PC can be synthesized directly from PG and CO2, the process can be regarded as DMC synthesis indirectly from MeOH and CO2 (Scheme 2b). In this context, the synthesis of cyclic carbonates from CO2 with glycol could be promising but remains in its infancy. We have already developed efficient metal catalysis for PC synthesis from PG and CO2, such as Bu2SnO or Bu2Sn(OMe)2,16 magnesium and its oxide.9 Despite high selectivity, PG conversion and PC yield were unsatisfactory [e.g. 3.4% yield with Bu2Sn(OMe)2] even under harsh reaction conditions (180 °C, 15 MPa). Such low conversion and yield are likely ascribed to the thermodynamic limitations and/or deactivation of the catalysts by the co-product water as shown in Scheme 3. Therefore, removing water to shifting the equilibrium to PC is the key to accomplishing high conversion. Various dehydrating agents have been introduced for this purpose. Inorganic dehydrating agents such as molecular sieve, CaCl2 are easily recyclable and do not form by-products. However, PC yields were still low because of the dehydration reversibility under the reaction conditions. In other words, it is difficult to physically absorb water at high temperature.


image file: c6ra04422k-s3.tif
Scheme 3 PC synthesis from PG and CO2 using PO as dehydrating agent.

As a complement, irreversible hydration with organic compounds is more effective at high temperature. For example, the hydration of nitriles to the amides can conveniently and selectively consume water.17–20 Therefore, acetonitrile has been applied as dehydrating agent to enhance the reaction between PG and CO2 with catalysts CeO2–ZrO2,21,22 alkali carbonate (i.e. Cs2CO3),23 modified zinc oxide (i.e. ZnO/KI),24 zinc acetate,25 and organic base (i.e. 1,5,7-triazabicyclo[4.4.0]dec-5-ene, TBD)26 respectively. The drawbacks associated with the use of chemical dehydrating agents are the high cost and the formation of hydrolytic products, which are hard to separate and recycle. The acetamide generated via hydrolysis of acetonitrile could further react with water to afford acetic acid which then reacts with PG to produce propylene glycol-2-acetate and thereby reducing the selectivity of PC.

Herein, we performed the K2CO3-catalyzed PC synthesis from PG and CO2 in PEG/CO2 biphasic system (Scheme 3). In this work, propylene oxide (PO) works as a dehydrating agent. The hydration of PO consumes water formed, and generates more PG, thus allows the equilibrium to shift to PC formation. The high efficiency may be explained by PEG coordinating with the potassium cation to enhance the basicity and nucleophilicity of carbonate anion.27,28 In addition, good solubility of CO2 and K2CO3 in PEG may also contribute to such promotive effect.32,33 As a consequence, this PEG/CO2 biphasic system by employing PO as a dehydrant could efficiently prepare PC with no by-product from the use of PO.

Results and discussion

In order to directly and efficiently prepare PC from PG and CO2, various bases including Cs2CO3, K2CO3, Bu2SnO, KF/Al2O3, CsOH and KOH were screened as the catalysts at 120 °C, 8 MPa. To our delight, when Cs2CO3 and K2CO3 used as catalysts with PO as dehydrating agent, PC was obtained in 19% and 12% yield, respectively, being consistent with the basicity (entry 1, 2, Table 1), while other catalysts gave only trace amounts of PC (entries 3–6). Control experiments were conducted to gain deeper insight into the role of PO in PC synthesis. Without PG or PO, almost no PC was detected by GC (entry 7, 8), suggesting that PG and PO are inevitable for PC formation. Either PG or PO alone could not react with CO2 to give PC under otherwise identical reaction conditions in this study. To distinguish PC produced from PG or PO, ethylene epoxide (EO) was chosen as a dehydrating agent because EC and PC can be unequivocally identified by GC technique. As a result, 24% of PC and 4% of EC were detected, indicating that the PC came mainly from PG rather than PO (entry 9). Herein, each PC yield in this study was obtained from two control experiments: a normal one to get the total PC yield from both PG and PO, another without PG to calculate the PC yield directly from PO, then PC yield from PG was obtained by subtracting such amount of PC from PO.
Table 1 PC synthesis from PG and CO2 using epoxide as dehydrating agenta

image file: c6ra04422k-u1.tif

Entry Catal. Dehydrating agent (2 eq.) Selectivityb (%) Yieldb (%)
a Reaction conditions: PG (2.5 mmol, 0.1902 g), catalyst (5 mol%), dehydrating agent (200 mol%, 5 mmol), reaction pressure (8 MPa), 120 °C, 12 h. b Determined by GC using biphenyl as internal standard. c Without PG. d GC yield of EC.
1 Cs2CO3 PO >99 19
2 K2CO3 PO >99 12
3 Bu3SnO PO >99 <1
4 KF/Al2O3 PO >99 3
5 CsOH PO >99 3
6 KOH PO >99 <1
7c Cs2CO3 PO >99 <1
8 Cs2CO3 >99 <1
9 Cs2CO3 EO >99 24 (4d)


As an inexpensive, non-toxic, nearly non-volatile and environmentally benign reaction medium,29–31 PEG is expandable with CO2,32,33 thus leading to changes in its physical properties, such as lowered melting points34 and lowered viscosity.35 When PEG was selected as solvent, the PC yield directly from PG distinctly increased to 55% after subtracting the PC amount formed from PO under the same conditions (entry 2, Table 2), indicating the introduction of PEG is beneficial to the carboxylation of PG and CO2 to produce PC when using PO as dehydrating agent. Typically, EO as a dehydrant afforded 44% yield of PC accompanying with 0.70 mmol of EC (entry 3). Accordingly, an epoxide could act as an efficient dehydrating agent to promote the carboxylation of PG with CO2 to produce PC in PEG/CO2 biphasic system. Notably, no additional by-product was detected in this process. Subsequently, the influence of PEG with different molecular weight ranging from 400 to 6000 was explored. When the reaction was conducted in PEG400, PO was quantitatively converted to PC but only 7% yield of PC came from PG (entry 3, Table 2). As PEG molecular weight increased to 800, PC from PG was increased to 66% (entry 4, 5). However, further increasing PEG molecular weight to 6000 leaded to a decrease in PC yield (entry 6), presumably being ascribed to the decreasing in coordination effect and CO2 solubility, as well as the increase in viscosity as PEG molecular weight increasing. Therefore, PEG800 was chosen as the appropriate solvent for further investigation.

Table 2 PC synthesis from PG in PEG/CO2 with epoxide as dehydranta

image file: c6ra04422k-u2.tif

Entry PEG PEG (g) PC amountb (mmol) PC yield from PGc (%)
a Reaction conditions: PG (2.5 mmol, 0.1902 g), Cs2CO3 (5 mol%, 0.0407 g), PO (5 mmol, 0.2904 g), reaction pressure (8 MPa), 120 °C, 12 h. b GC yield with biphenyl as internal standard. c Obtained by subtracting the PC yield from PO under the same conditions. d Without PG.
1d PEG1000 0.5 0.44
2 PEG1000 0.5 1.86 55
3 PEG400 0.2 0.10 7
4 PEG600 0.3 1.59 54
5 PEG800 0.4 1.27 66
6 PEG6000 0.5 1.08 44


Then, the influence of different basic catalysts was examined in PEG800, as shown in Table 3. Without any catalyst, no PC directly from PG was detected (entry 1). Alkali carbonates e.g. Cs2CO3 and K2CO3 gave relatively higher yields (entry 2, 3); whereas Na2CO3 was less effective (entry 4). In particular, Cs2CO3 showed lower activity than K2CO3 due to favourable coordination between PEG800 and K+. Others, such as CaCO3, MgO, CaO did not work at all under the given conditions (entries 5–7). Therefore, K2CO3 was proved to be the most appropriate catalyst for PC synthesis. Furthermore, increasing PEG800 amount gave no significant rise in PC yield (entries 8–10). Hence, 0.4 g of PEG800 was chosen as reaction medium. In addition, changing the reaction temperature to 100 °C or 140 °C resulted in adverse effect (entry 9, 10), presumably being ascribed to intermolecular dehydrating condensation of PG and PEG.28

Table 3 Influence of catalyst and PEG amount for PC synthesis from PG and CO2a
Entry Catalyst PEG800 (g) Yieldb,c (%)
a Reaction conditions: PG (2.5 mmol, 0.1902 g), PO (2 equiv. relative to PG, 0.2904 g), catalyst (5 mol%), in PEG800 under 8 MPa pressure, 120 °C for 12 h. b Determined by GC using biphenyl as internal standard. c PC yield relative to PG. d 100 °C. e 140 °C.
1 0.4 <1
2 Cs2CO3 0.4 66
3 K2CO3 0.4 74
4 Na2CO3 0.4 15
5 CaCO3 0.4 <1
6 MgO 0.4 <1
7 CaO 0.4 <1
8 K2CO3 0.5 75
9d K2CO3 0.4 38
10e K2CO3 0.4 <1


With those results in hand, we then investigated the effect of catalyst loading, reaction pressure and time on the reaction outcome as listed in Table 4. Reducing K2CO3 loading from 5 to 3 mol% dramatically leaded to decrease in PC yield (entry 2 vs. 1). However, when increased the catalyst amount to 10 mol%, no beneficial effect was shown (entry 3).

Table 4 Influence of catalyst loading, pressure and reaction timea
Entry K2CO3 (mol%) Pressure (MPa) Time (h) Yieldb,c (%)
a Reaction conditions: PG (2.5 mmol, 0.1902 g), PO (2 equiv. to PG, 0.2904 g), in 0.4 g PEG800 with CO2 and K2CO3, 120 °C. b Determined by GC using biphenyl as internal standard. c Yield of PC relative to PG.
1 3 8 12 55
2 5 8 12 74
3 10 8 12 68
4 5 2 12 49
5 5 4 12 79
6 5 10 12 66
7 5 4 10 78
8 5 4 6 70


Furthermore, the dependence of PC yield on reaction pressure was carefully examined. PC yield gradually increased when raising reaction pressure from 2 to 4 MPa (entry 4, 5), being ascribed to CO2 amount increase. However, too high reaction pressure would dilute the reactants concentration,28 leading to a decrease in PC yield (entry 6). The effect of reaction time on PC synthesis was also explored. 78% and 70% yield of PC were obtained when the reactions stopped at 10 h and 6 h, respectively (entry 7, 8). Consequently, the most suitable reaction conditions for the carboxylation of PG with CO2 using PO as dehydrate agent were 5 mol% K2CO3 in 0.4 g PEG800 under 4 MPa pressure at 120 °C for 10 h.

The proposed mechanism for the present PC synthesis from PG and CO2 in PEG/CO2 biphasic system catalyzed by K2CO3 with PO as dehydrating agent was depicted in Scheme 4. The solubility of CO2 and K2CO3 in PEG is excellent, and PEG can coordinate with K+ to enhance the basicity and nucleophilicity of the carbonate anion.27,28,32,33 The deprotonation of PG by K2CO3 and subsequent insertion of CO2 gives the carbonate intermediate.


image file: c6ra04422k-s4.tif
Scheme 4 Proposed mechanism for PC formation from PG and CO2 by using PO as dehydrating agent.

Then PC is generated along with the formation of an equivalent water. The water further reacts with PO, regenerating more PG; and thus both shifts the equilibrium to PC production. After the completion of the carboxylation of PG with CO2, PC could be separated or subsequently reacts with methanol to produce DMC through transesterification reaction. Another reaction in this system is the direct cyclization of PO and CO2 to afford PC.

Conclusions

In conclusion, we have developed a simple, ecologically safe, cost-effective and industrially feasible process for PC synthesis from PG with CO2 using K2CO3 as catalyst in PEG800/CO2 biphasic system. 78% yield of PC from PG can be attained under mild reaction conditions (120 °C, 4 MPa). In this process, PO is utilized as dehydrating agent to react with co-produced water along with regenerating PG, both are favorable to shift the equilibrium to PC formation. Compared with common organic dehydrating agents, no additional by-product was generated in this process, resulting in workup procedure and economic benefits.

Experimental

General information

Alcohols, epoxides and PEGs were purchased from Aladdin. PG was dried with anhydrous Na2SO4, then distilled under reduced pressure prior to use. PEGs were dried in vacuum oven at 80 °C for two days and all the catalysts were dried in vacuo at 150 °C for 4 h. CO2 was commercially available with a purity of 99.99%. Other reagents including diethyl ether, ethyl acetate, petroleum ether (60–90 °C) were purchased from Tianjin Guangfu Fine Chemical Research Institute without further purification. GC analyses were performed on Shimadzu GC-2014 equipped with capillary column (RTX-17 30 m × 0.25 μm) and flame ionization detector. GC-MS analyses were performed on Shimadzu 2014-QP2010 SE. NMR spectra were recorded on Bruker 400 in CDCl3. 1H and 13C NMR chemical shifts (δ) were given in ppm relative to CDCl3 (7.26 ppm and 77.0 ppm).

General procedure for the K2CO3-promoted PC synthesis from PG and CO2

A mixture of PG (2.5 mmol, 0.1902 g), potassium carbonate (17.3 mg, 0.125 mmol, 5 mol% relative to PG), epoxide (5 mmol) and PEG800 (0.5 mmol, 0.4 g) were placed in a 25 mL autoclave equipped with inner glass tube. 2 MPa CO2 was sequentially introduced into the autoclave and heated to 120 °C. The final pressure at the reaction temperature was regulated by introducing necessary amount of CO2 and the mixture was stirred continuously for 12 h. After the reaction, the reactor was cooled to 0 °C in ice water. Then CO2 was released slowly, the product was extracted with diethyl ether after depressurization. Then gas chromatograph analysis was performed with biphenyl as internal standard. The product was further identified using GC-MS by comparing retention times and fragmentation patterns with authentic samples and were also isolated by column chromatography on silica gel (200–300 mesh, petroleum ether (60–90 °C)/EtOAc = 3[thin space (1/6-em)]:[thin space (1/6-em)]1) and identified by NMR and MS measurement.

4-Methyl-1,3-dioxolan-2-one (PC)

Colorless oil; 1H NMR (400 MHz, CDCl3): δ = 4.93–4.77 (m, 1H, CH), 4.55 (t, J = 8.0 Hz, 1H, CH2), 4.02 (t, J = 7.8 Hz, 1H, CH2), 1.48 (d, J = 6.2 Hz, 3H, CH3); 13C NMR (101 MHz, CDCl3): δ = 155.0 (C), 73.5 (CH), 70.7 (CH2), 19.5 (CH3). EI-MS, m/z (%): 103.14 (26) [M]+, 57.03 (100).

1,3-Dioxolan-2-one (EC)

Colorless crystal; 1H NMR (400 MHz, CDCl3): δ = 4.50 (s, 4H, CH2); 13C NMR (101 MHz, CDCl3): δ = 155.5 (C), 64.6 (CH2).

Acknowledgements

This work was financially supported by the National Natural Sciences Foundation of China, the “111” Project of the Ministry of Education of China (No. B06005), Specialized Research Fund for the Doctoral Program of Higher Education (20130031110013), MOE Innovation Team (IRT13022) of China.

References

  1. T. Sakakura and K. Kohno, Chem. Commun., 2009, 1312–1330 RSC.
  2. B. Schäffner, J. Holz, S. P. Verevkin and A. Börner, ChemSusChem, 2008, 1, 249–253 CrossRef PubMed.
  3. J. Bayardon, J. Holz, B. Schäffner, V. Andrushko, S. Verevkin, A. Preetz and A. Börner, Angew. Chem., Int. Ed., 2007, 46, 5971–5974 CrossRef CAS PubMed.
  4. B. Schäffner, F. Schäffner, S. P. Verevkin and A. Börner, Chem. Rev., 2010, 110, 4554–4581 CrossRef PubMed.
  5. S. Fukuoka, M. Kawamura, K. Komiya, M. Tojo, H. Hachiya, K. Hasegawa, M. Aminaka, H. Okamoto, I. Fukawa and S. Konno, Green Chem., 2003, 5, 497–507 RSC.
  6. P. Tundo and M. Selva, Acc. Chem. Res., 2002, 35, 706–716 CrossRef CAS PubMed.
  7. M. A. Pacheco and C. L. Marshall, Energy Fuels, 1997, 11, 2–29 CrossRef CAS.
  8. Z.-Z. Yang, L.-N. He, X.-Y. Dou and S. Chanfreau, Tetrahedron Lett., 2010, 51, 2931–2934 CrossRef CAS.
  9. Y. Du, L.-N. He and D.-L. Kong, Catal. Commun., 2008, 9, 1754–1758 CrossRef CAS.
  10. D. Delledonne, F. Rivetti and U. Romano, Appl. Catal., A, 2001, 221, 241–251 CrossRef CAS.
  11. M. Honda, S. Kuno, S. Sonehara, K.-i. Fujimoto, K. Suzuki, Y. Nakagawa and K. Tomishige, ChemCatChem, 2011, 3, 365–370 CrossRef CAS.
  12. M. Honda, M. Tamura, Y. Nakagawa, S. Sonehara, K. Suzuki, K.-i. Fujimoto and K. Tomishige, ChemSusChem, 2013, 6, 1341–1344 CrossRef CAS PubMed.
  13. J.-S. Tian, C.-X. Miao, J.-Q. Wang, F. Cai, Y. Du, Y. Zhao and L.-N. He, Green Chem., 2007, 9, 566–571 RSC.
  14. M. He, Y. Sun and B. Han, Angew. Chem., Int. Ed., 2013, 52, 9620–9633 CrossRef CAS PubMed.
  15. Z.-Z. Yang, Y.-N. Zhao, L.-N. He, J. Gao and Z.-S. Yin, Green Chem., 2012, 14, 519–527 RSC.
  16. Y. Du, D.-L. Kong, H.-Y. Wang, F. Cai, J.-S. Tian, J.-Q. Wang and L.-N. He, J. Mol. Catal. A: Chem., 2005, 241, 233–237 CrossRef CAS.
  17. T. Sakakura, J.-C. Choi and H. Yasuda, Chem. Rev., 2007, 107, 2365–2387 CrossRef CAS PubMed.
  18. K. Yamaguchi, M. Matsushita and N. Mizuno, Angew. Chem., Int. Ed., 2004, 116, 1602–1606 CrossRef.
  19. J. N. Moorthy and N. Singhal, J. Org. Chem., 2005, 70, 1926–1929 CrossRef CAS PubMed.
  20. E. Da Silva, W. Dayoub, G. Mignani, Y. Raoul and M. Lemaire, Catal. Commun., 2012, 29, 58–62 CrossRef CAS.
  21. K. Tomishige, H. Yasuda, Y. Yoshida and M. Nurunnabi, Catal. Lett., 2004, 95, 45–49 CrossRef CAS.
  22. K. Tomishige, H. Yasuda, Y. Yoshida, M. Nurunnabi, B. Li and K. Kunimori, Green Chem., 2004, 6, 206–214 RSC.
  23. S. Huang, S. Liu, J. Li, N. Zhao, W. Wei and Y. Sun, Catal. Lett., 2006, 112, 187–191 CrossRef CAS.
  24. S. Huang, S. Liu, J. Li, N. Zhao, W. Wei and Y. Sun, Catal. Lett., 2007, 118, 290–294 CrossRef CAS.
  25. X. Zhao, N. Sun, S. Wang, F. Li and Y. Wang, Ind. Eng. Chem. Res., 2008, 47, 1365–1369 CrossRef CAS.
  26. S. Huang, J. Ma, J. Li, N. Zhao, W. Wei and Y. Sun, Catal. Commun., 2008, 9, 276–280 CrossRef CAS.
  27. A.-H. Liu, R. Ma, C. Song, Z.-Z. Yang, A. Yu, Y. Cai, L.-N. He, Y. N. Zhao, B. Yu and Q.-W. Song, Angew. Chem., Int. Ed., 2012, 51, 11306–11310 CrossRef CAS PubMed.
  28. D.-L. Kong, L.-N. He and J.-Q. Wang, Synlett, 2010, 1276–1280 CAS.
  29. Z. Hou, N. Theyssen, A. Brinkmann and W. Leitner, Angew. Chem., Int. Ed., 2005, 44, 1346–1349 CrossRef CAS PubMed.
  30. J. Q. Wang, F. Cai, E. Wang and L. N. He, Green Chem., 2007, 9, 882–887 RSC.
  31. Y. Du, Y. Wu, A.-H. Liu and L.-N. He, J. Org. Chem., 2008, 73, 4709–4712 CrossRef CAS PubMed.
  32. Z.-M. Xue, J.-L. Zhang, L. Peng, J.-S. Li, T.-C. Mu, B.-X. Han and G.-Y. Yang, Angew. Chem., Int. Ed., 2012, 51, 12325–12329 CrossRef CAS PubMed.
  33. R.-C. Luo, X.-T. Zhou, S.-Y. Chen, Y. Li, L. Zhou and H.-B. Ji, Green Chem., 2014, 16, 1496–1506 RSC.
  34. D. J. Heldebrant and P. G. Jessop, J. Am. Chem. Soc., 2003, 125, 5600–5601 CrossRef CAS PubMed.
  35. S. Yeo and E. Kiran, J. Appl. Polym. Sci., 2000, 75, 306–315 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra04422k

This journal is © The Royal Society of Chemistry 2016