Next Article in Journal
Selective Oxidation of Raw Glycerol Using Supported AuPd Nanoparticles
Next Article in Special Issue
Magnetically Separable and Sustainable Nanostructured Catalysts for Heterogeneous Reduction of Nitroaromatics
Previous Article in Journal / Special Issue
Application of “Boomerang” Linear Polystyrene-Stabilized Pd Nanoparticles to a Series of C-C Coupling Reactions in Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Guanidine Hydrochloride/ZnI2 as Heterogeneous Catalyst for Conversion of CO2 and Epoxides to Cyclic Carbonates under Mild Conditions

1
The Academy of Fundamental and Interdisciplinary Science, Harbin Institute of Technology, Harbin 150080, China
2
School of Life Science and Technology, Harbin Institute of Technology, Harbin 150080, China
3
State Key Laboratory of Urban Water Resource and Environment, Harbin Institute of Technology, Harbin 150080, China
*
Author to whom correspondence should be addressed.
Catalysts 2015, 5(1), 119-130; https://doi.org/10.3390/catal5010119
Submission received: 24 December 2014 / Accepted: 9 February 2015 / Published: 11 February 2015
(This article belongs to the Special Issue Metal Catalysts Recycling and Heterogeneous/Homogeneous Catalysis)

Abstract

:
In this article, the combination of guanidine hydrochloride with co-catalyst ZnI2 proved to be a highly efficient heterogeneous catalyst for the environmentally benign, solvent-free synthesis of cyclic carbonates under mild reaction conditions. The effects of different co-catalysts as well as reaction parameters including catalyst loadings, CO2 pressure, reaction temperature, and reaction time on the coupling reaction of CO2 to propylene oxide were thoroughly investigated. With the molar ratio of guanidine hydrochloride to ZnI2 at 5:1, excellent yield (94%) and selectivity (≥99%) of propylene carbonate were obtained under 100 °C and at 1 MPa for 1.5 h. Additionally, ZnI2 could be recycled, but because of the washing loss of guanidine hydrochloride, there was a slight decrease in the yield of propylene carbonate. Gratifyingly, the activity of the catalytic system could be restored by adding additional 20 mol% of fresh guanidine hydrochloride, thus exhibiting excellent recyclability of the ZnI2 catalyst. Moreover, the binary catalysts were also versatile when using other epoxides for CO2 cycloaddition. A possible reaction mechanism was proposed wherein guanidine hydrochloride plays a dual role in activating CO2 and epoxide, and ZnI2 activated epoxide, simultaneously. The synergistic effect of guanidine hydrochloride and ZnI2 ensure the reaction proceeds effectively.

1. Introduction

Carbon dioxide can be specially spotlighted as an abundant, inexpensive, and renewable C1 resource, which can replace commonly used toxic C1 building blocks such as phosgene [1]; however at the same time CO2 is also the main greenhouse gas contributor [2]. Therefore, chemical fixation of CO2 has been a hot topic from the viewpoint of sustainable development and green chemistry. In this regard, one of the most promising strategies is the transformation of CO2 with epoxides to yield the corresponding cyclic carbonates (Scheme 1) [3], which is considered green because of 100% atom efficiency. Also cyclic carbonates can be widely applied in proton inert solvents, as precursors for polycarbonates and polyurethanes synthesis, as electrolytes in lithium secondary batteries, and as intermediates for production of pharmaceuticals, fine chemicals and agricultural chemicals [4,5].
Scheme 1. Cycloaddition of epoxide with CO2.
Scheme 1. Cycloaddition of epoxide with CO2.
Catalysts 05 00119 g005
So far, numerous homogeneous and heterogeneous catalysts for this transformation have been reported, including metal-oxides [6,7], alkali metal salts [8], organometallic complexes [9], organic bases or ammonium/phosphonium salts [10,11,12], N-heterocyclic carbenes [13], modified molecular sieves [14,15] and ionic liquids [16,17,18]. Nevertheless, homogeneous catalysts are undesirable in industry due to problems such as rigorous separation and purification of the products. Also the developed heterogeneous catalysts are somewhat restricted due to low activity, harsh reaction conditions, or poor stability [19]. Therefore, it is of still great importance to explore heterogeneous catalysts with excellent activity and stability under mild reaction conditions for CO2 cycloaddition.
In order to improve the catalytic performance, the catalysts need to simultaneously possess Lewis acid and base to activate epoxides and CO2, respectively. Additionally, it has been reported that zinc halides can coordinate with the oxygen atom to activate the epoxide by forming the zinc-epoxide adduct [20,21]. Moreover, theoretical study also indicated that the formation of a hydrogen bond between catalyst and epoxide could accelerate the ring-opening reaction of the epoxide, thus enhancing the activity significantly [22]. Guanidine hydrochloride (GndCl) is a cheap and nontoxic chemical that has two primary amino nucleophilic groups and a chlorid ion, and can serve as activator of CO2 as well as epoxide and attacking reagent simultaneously, but still has never been used in catalytic CO2 cycloaddition with epoxides. Herein, we developed a dual heterogeneous catalyst system composed of guanidine hydrochloride and ZnI2 for the conversion of CO2 and epoxides to cyclic carbonates. The effects of various Lewis acid co-catalysts as well as reaction parameters such as CO2 pressure, catalyst loadings, reaction temperature, and reaction time on catalytic activity were investigated thoroughly. GndCl/ZnI2 was the most efficient among these developed binary catalysts and showed an excellent synergetic effect in promoting the reaction under mild conditions. Moreover, the reusability of the catalytic system was investigated and a possible synergistic catalytic mechanism was proposed.

2. Results and Discussion

2.1. Screening of Catalysts and Their Reusability

The catalytic performance of GndCl with different Lewis acid co-catalysts was investigated with the coupling reaction of CO2 to propylene oxide (PO), and the results are shown in Table 1. Almost no activity in forming the product propylene carbonate (PC) was observed with GndCl or ZnI2 alone (Entries 1 and 2) and only 27% yield of PC was obtained when GndCl was employed under 130 °C and at 3.0 MPa CO2 for 1.5 h (Entry 3), thus indicating the bifunctional role of GndCl to catalyze the cycloaddition reaction. Noticeably, the combination of GndCl with ZnI2 exhibited remarkably improved catalytic activity, and PC yield approached 94% at 100 °C and 1.0 MPa CO2 for 1.5 h (Entry 4) without any solvent. It was noted that the selectivity to PC was always above 99%, the rest was due to the trace amounts of PO polymerization in side reactions. Additionally, the performances of various Lewis acid co-catalysts for this transformation were explored. In comparison, almost no product was detected when ZnO or Zn(NO3)2 was used (Entries 7 and 8) and the reaction hardly occurred or very low conversions were obtained over GndCl with Cd2+, Mn2+, Co2+ or Fe3+ (Entries 9–12). The catalytic activity improved with the stronger anionic nucleophilic property of zinc halide in the order of ZnI2 >>ZnBr2 >ZnCl2, which was in accordance with previous works [8,23]. When ZnBr2 and ZnCl2 were used as the co-catalysts, 43% and 12% yield of PC were obtained, respectively (Entries 5 and 6), and ZnI2 co-catalyst displayed a remarkable role in accelerating the reaction with 94% PC yield. GndCl combined with ZnI2 showed higher activity than other metallic halides. Hence, GndCl coordinated with ZnI2 was chosen for further investigation.
In the meanwhile, the GndCl dosage was a key factor to the catalytic activity. 0.5 mmol and 1.0 mmol GndCl gave 37% and 94% yields of PC (Entries 4 and 13) and further improvement of GndCl dosage at 1.5 mmol resulted in a slight increase at 95% (Entry 14). Hence, 1.0 mmol GndCl was adopted for further investigation.
For potential application in a green and economic process, the catalyst recyclability plays an important role. Experiments were carried out to test the recyclability of GndCl/ZnI2 at 100 °C and 1.0 MPa CO2, the results are shown in Table 1. There was a slight decrease in PC yield in the second run (Entry 15) because of the washing loss of catalysts. After an additional 20 mol% of fresh guanidine hydrochloride was added together with recyclable ZnI2 and GndCl for the third run, the PC yield again reached 90% (Entry 16).
Table 1. Screening of catalysts for the cycloaddition reaction a.
Table 1. Screening of catalysts for the cycloaddition reaction a.
EntryCatalyst systemCatalytic results b
Yield (%)Selectivity (%)
1GndCltrace
2ZnI2trace
3 cGndCl27≥98
4GndClZnI294≥99
5GndClZnBr243≥99
6GndClZnCl212≥99
7GndClZnOtrace≥99
8GndClZn(NO3)2trace
9GndClCdCl2trace
10GndClMnCl2trace
11GndClCoCl2trace
12GndClFeCl35≥99
13 dGndClZnI237≥99
14 eGndClZnI295≥99
15 fGndClZnI280≥99
16 gGndClZnI290≥98
a Reaction conditions: PO 34.5 mmol, GndCl 1.0 mmol, co-catalyst 0.2 mmol, 100 °C, CO2 pressure 1.0 MPa, 1.5 h. b Determined by GC. c CO2 pressure 3.0 MPa, 130 °C, 1.5 h. d GndCl 0.5 mmol. e GndCl 1.5 mmol. f The second run. g The third run, additional 20 mol% of GndCl was added.

2.2. Effect of Reaction Parameters

To investigate the influence of catalyst loading on the PC synthesis, a series of experiments were carried out by varying the molar ratio of ZnI2 catalyst to propylene oxide, and the results are shown in Figure 1. With the ZnI2/PO molar ratio increased from 0.29–0.58%, there was a sharp increase of PC yield from 46%–94%. When ZnI2 loading was further increased to 0.87 mol%, PC yield did not apparently rise. Even at a low loading of ZnI2 at 0.58 mol%, high PC yield (94%) and selectivity (≥99%) were obtained. Gratifyingly, the PC selectivities hardly changed on increasing the ZnI2 catalyst loading from 0.29–0.87 mol%. Therefore, 0.58 mol% was chosen as the optimal amount of ZnI2.
Generally, a significant drawback associated with CO2 as reagent or reaction medium in organic synthesis, is the potential danger of operating at high pressure [24]. Hence, the effect of CO2 pressure on the CO2 coupling reaction was investigated. As shown in Figure 2, a dramatic increase of PC yield was detected in the low-pressure range of 0.5–1.0 MPa. Also, the yields remained nearly constant at medium CO2 pressures of 1.0–3.0 MPa, suggesting that CO2 was not involved in the rate-determining step in this pressure range [25]. The desired product yield was obtained even at low CO2 pressure of 1.0 MPa over GndCl/ZnI2 catalysts. A similar effect of CO2 pressure on catalytic activity was observed in other catalytic systems [6,25,26,27], which could be explained by the fact that the introduced CO2 was dissolved in propylene oxide or “liquefied” by the formation of a CO2-PO complex [23]. In the low-pressure range, increasing CO2 pressure led to an increase of CO2-PO concentration in the liquid phase and thus enhanced the PC yield. However, on increasing CO2 pressure, the concentration of PO decreased in the liquid phase and increased in the vapor phase, which gave rise to the constant PC yield. Thus, 1.0 MPa was chosen as the optimal CO2 pressure.
Figure 1. Influence of ZnI2 loading on propylene carbonate (PC) yield and selectivity. Reaction conditions: propylene oxide (PO) 34.5 mmol, GndCl 1.0 mmol, CO2 pressure 1.0 MPa, 100 °C, 1.5 h.
Figure 1. Influence of ZnI2 loading on propylene carbonate (PC) yield and selectivity. Reaction conditions: propylene oxide (PO) 34.5 mmol, GndCl 1.0 mmol, CO2 pressure 1.0 MPa, 100 °C, 1.5 h.
Catalysts 05 00119 g001
Figure 2. Influence of CO2 pressure on the cycloaddition reaction. Reaction conditions: PO 34.5 mmol, GndCl 1.0 mmol, ZnI2 0.2 mmol, 100 °C, 1.5 h.
Figure 2. Influence of CO2 pressure on the cycloaddition reaction. Reaction conditions: PO 34.5 mmol, GndCl 1.0 mmol, ZnI2 0.2 mmol, 100 °C, 1.5 h.
Catalysts 05 00119 g002
The influence of reaction temperature on PC yield and selectivity is depicted in Figure 3. It is obvious that the reaction was sensitive to reaction temperature. With the temperature increasing from 80–100 °C, PC yield rapidly increased from 41%–94% and PC selectivity stayed invariably at 99%. As the temperature was further raised to 120 °C, PC yield increased slightly to 98%. Hence, 100 °C was the optimal reaction temperature for PC synthesis in the light of energy consumption and PC yield.
Figure 3. Influence of reaction temperature on the cycloaddition reaction. Reaction conditions: PO 34.5 mmol, GndCl 1.0 mmol, ZnI2 0.2 mmol, CO2 pressure 1.0 MPa, 1.5 h.
Figure 3. Influence of reaction temperature on the cycloaddition reaction. Reaction conditions: PO 34.5 mmol, GndCl 1.0 mmol, ZnI2 0.2 mmol, CO2 pressure 1.0 MPa, 1.5 h.
Catalysts 05 00119 g003
The effect of reaction time on PC synthesis is illustrated in Figure 4. When the cycloaddition reaction proceeded for 1.5 h, PC yield rapidly achieved 94%. Afterwards, there appeared a gradual increase of PC yield. Moreover, the selectivity to PC stayed above 98% throughout.
Figure 4. Influence of reaction time on the cycloaddition reaction. Reaction conditions: PO 34.5 mmol, GndCl 1.0 mmol, ZnI2 0.2 mmol, CO2 pressure 1.0 MPa, 100 °C.
Figure 4. Influence of reaction time on the cycloaddition reaction. Reaction conditions: PO 34.5 mmol, GndCl 1.0 mmol, ZnI2 0.2 mmol, CO2 pressure 1.0 MPa, 100 °C.
Catalysts 05 00119 g004

2.3. Catalytic Activity towards Other Epoxides

In order to study the efficiency and general applicability of the developed GndCl/ZnI2 catalytic system, the cycloadditions of CO2 with different substituent epoxides were conducted, and the results are exhibited in Table 2. GndCl/ZnI2 displayed efficient activities for the conversions of a variety of epoxides to the corresponding cyclic carbonates with good to excellent yields and high selectivities. PO was the most active of the epoxides. Also, various epoxides bearing electron-donating and electron-withdrawing groups provided the desired products in appreciably shorter time (Entries 1–3), thus reflecting the outstanding efficiency of GndCl/ZnI2 catalysts. However, with the increase of steric hindrance from side chains of substrates, the catalytic activities of CO2 cycloaddition to styrene oxide and cyclohexene oxide were impeded, however, satisfactory yields were also obtained at the expense of longer reaction time or higher reaction temperature compared with the terminal epoxides (Entries 4 and 5).
Table 2. Catalytic activity of CO2 cycloaddition to various epoxides a.
Table 2. Catalytic activity of CO2 cycloaddition to various epoxides a.
EntryEpoxideProductTime (h)Reaction results b
Yield (%)Selectivity (%)
1 c Catalysts 05 00119 i001 Catalysts 05 00119 i0061.59499
2 Catalysts 05 00119 i002 Catalysts 05 00119 i0071.59299
3 Catalysts 05 00119 i003 Catalysts 05 00119 i0081.59299
4 Catalysts 05 00119 i004 Catalysts 05 00119 i00938098
5 d Catalysts 05 00119 i005 Catalysts 05 00119 i010117198
a Reaction conditions: epoxide 13.9 mmol, GndCl 1.0 mmol, ZnI2 0.2 mmol, CO2 pressure 1.0 MPa, 100 °C. b Determined by GC. c PO 34.5 mmol. d Reaction temperature 120 °C.

2.4. Proposed Reaction Mechanism

In previous works [6,19], it was regarded that the hydrogen bond formed between the hydrogen atom of the amine and the oxygen atom of epoxide would lead to polarization of the C-O bond, which could then facilitate the ring opening of the epoxide; and simultaneously, the primary amino group of GndCl would activate CO2 to form a carbamate salt (Scheme 2) [19,28,29], which would facilitate the CO2 cycloaddition reaction. Moreover, it was recognized that epoxides could be activated by zinc halides by the formation of a zinc-epoxide adduct [30], which had a positive effect on the ring opening of the epoxide. Based on this, a plausible mechanism is proposed as shown in Scheme 3. Firstly, GndCl plays the role of Lewis base to activate CO2 to form a zwitterionic adduct (intermediate 2). Meanwhile, the epoxide is activated by the formation of a hydrogen bond between the hydrogen atom of the amino and oxygen atoms of the epoxide; additionally, the epoxide is further activated by the zinc atom from ZnI2 forming a Zn-epoxide adduct. Both the integrated activations from GndCl and ZnI2 lead to easy polarization of the C-O bond of the epoxide. Secondly, the I anion from ZnI2 plays the role of Lewis base to attack the less sterically hindered β-carbon atom of the epoxide, assisting the ring-opening with the formation of the intermediate 1. Subsequently, the GndCl simultaneously coordinates with CO2, affording a carbamate salt 2 as the activated species of CO2 [29,31]. Thereafter, the carbamate salt 2 undergoes electrophilic attack on the intermediate 1 to produce the new alkyl carbonate compound 3, simultaneously leaving one GndCl. Eventually, by the ensuing intramolecular ring-closure, the cyclic carbonate is formed and the catalysts regenerated. The synergetic roles played by GndCl and ZnI2 ensure the reaction proceeds effectively. Also, the catalytic activity of ZnI2 is higher than ZnBr2 and ZnCl2, which can well be explained by the nucleophilic property of the anion.
Scheme 2. Structure and formation of carbamate salt 2.
Scheme 2. Structure and formation of carbamate salt 2.
Catalysts 05 00119 g006
Scheme 3. Possible reaction mechanism for GndCl/ZnI2 catalyzed cycloaddition reaction.
Scheme 3. Possible reaction mechanism for GndCl/ZnI2 catalyzed cycloaddition reaction.
Catalysts 05 00119 g007

3. Experimental Section

3.1. Materials and Instruments

The reagents and chemicals (analytic grade unless otherwise stated) were purchased from Sinopharm Chemical Reagent Co., Ltd. (No. 52 Ningbo Road, Shanghai, China). Carbon dioxide with 99.99% purity was purchased from Harbin Qinghua Industrial gases Co., Ltd. (No. 46 Hongli Street, Xiangfang District, Harbin, China). Propylene oxide (99%), 1, 2-butylene oxide, epichlorhydrin, styrene oxide, cyclohexene oxide were purchased from Beijing InnoChem Science & Technology Co., Ltd. (No. 170 Beiyuan Road, Chaoyang District, Beijing, China) Guanidine hydrochloride (99%), ZnI2 (≥98%) were purchased from Adamas Reagent Co., Ltd. (No. 245 Jiachuan Road, Xuhui District, Shanghai, China).
GC analyses were performed on an Agilent GC-7890A equipped with a capillary column (Agilent 19091J-413).

3.2. Typical Procedure for the CO2 Cycloaddition Reaction to Propylene Oxide

The cycloaddition reaction of propylene oxide (PO) and CO2 was conducted in a 50 mL high pressure stainless-steel autoclave equipped with a magnetic stirring bar. In a typical run, a certain amount of GndCl/ZnI2 catalyst and PO (34.5 mmol) were added into the reactor successively, then the reactor was heated using an oil bath. At a fixed reaction temperature, CO2 was introduced into the reactor to a fixed pressure, and then the autoclave was heated at that temperature for a desired time. After the reaction was completed, the reactor was cooled to 0 °C in ice-water bath, and the remaining CO2 was released slowly to avoid the rest of the PO escaping with the gas. The catalysts were separated from the system by centrifugation, then washed with toluene (3 × 5 mL), dried and reused for the next run. The products were analyzed by GC with a flame ionization detector.

4. Conclusions

In summary, we developed a cheap, readily available and highly active GndCl/ZnI2 catalytic system for the synthesis of cyclic carbonates under mild, solvent-free conditions. GndCl/ZnI2 catalysts proceeded efficiently for the CO2 cycloaddition with different substituted epoxides, even with the less active epoxides such as styrene oxide and cyclohexene oxide, due to the synergetic effects of guanidine hydrochloride and ZnI2. The catalysts could be reused at least three times, however at the expense of an additional 20 mol% fresh guanidine hydrochloride to compensate for the partial loss on washing during recycles. Compared with previous heterogeneous catalysts, the catalytic system developed in this study was cheap, easily available, solvent-free for CO2 cycloaddition and under mild reaction conditions, which is of benefit for practical applications of CO2 conversion.

Acknowledgments

We sincerely acknowledge the financial supports from the National Natural Science Foundation of China (21373069), the Science Foundation of Harbin City (NJ20140037), the State Key Lab of Urban Water Resource and Environment of Harbin Institute of Technology (HIT2015DX08) and the Fundamental Research Funds for the Central Universities (HIT. IBRSEM. 201327).

Author Contributions

Bo Liu and Jianmin Sun contributed to the experiment design, the article writing and revising. Bo Liu also contributed to all the experimental data collection, and Mengshuai Liu and Lin Liang contributed to product analyses as well as helpful discussion.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aresta, M.; Dibenedetto, A. Utilisation of CO2 as a chemical feedstock: Opportunities and challenges. Dalton Trans. 2007, 2975–2992. [Google Scholar]
  2. Lee, T.; van de Meene, S. Comparative studies of urban climate co-benefits in Asian cities: An analysis of relationships between CO2 emissions and environmental indicators. J. Clean. Prod. 2013, 58, 15–24. [Google Scholar]
  3. Decortes, A.; Castilla, A.M.; Kleij, A.W. Salen-complex-mediated formation of cyclic carbonates by cycloaddition of CO2 to epoxides. Angew. Chem. Int. Ed. 2010, 49, 9822–9837. [Google Scholar]
  4. Sakakura, T.; Choi, J.C.; Yasuda, H. Transformation of carbon dioxide. Chem. Rev. 2007, 107, 2365–2387. [Google Scholar]
  5. Kuhl, K.P.; Cave, E.R.; Abram, D.N.; Jaramillo, T.F. New insights into the electrochemical reduction of carbon dioxide on metallic copper surfaces. Energy Environ. Sci. 2012, 5, 7050–7059. [Google Scholar]
  6. Yamaguchi, K.; Ebitani, K.; Yoshida, T.; Yoshida, H.; Kaneda, K. Mg-Al mixed oxides as highly active acid-base catalysts for cycloaddition of carbon dioxide to epoxides. J. Am. Chem. Soc. 1999, 121, 4526–4527. [Google Scholar]
  7. Dai, W.L.; Yin, S.F.; Guo, R.; Luo, S.L.; Du, X.; Au, C.T. Synthesis of propylene carbonate from carbon dioxide and propylene oxide using Zn-Mg-Al composite oxide as high-efficiency catalyst. Catal. Lett. 2010, 136, 35–44. [Google Scholar]
  8. Sako, T.; Fukai, T.; Sahashi, R. Cycloaddition of oxirane group with carbon dioxide in the supercritical homogeneous state. Ind. Eng. Chem. Res. 2002, 41, 5353–5358. [Google Scholar]
  9. Ahmadi, F.; Tangestaninejad, S.; Moghadam, M.; Mirkhani, V.; Baltork, I.M.; Khosropour, A.R. Electron-deficient tin (IV) tetraphenylporphyrin perchlorate: A highly efficient catalyst for chemical fixation of carbon dioxide. Polyhedron 2012, 32, 68–72. [Google Scholar]
  10. Sun, J.; Zhang, S.J.; Cheng, W.G.; Ren, J.Y. Hydroxyl-functionalized ionic liquid: A novel efficient catalyst for chemical fixation of CO2 to cyclic carbonate. Tetrahedron Lett. 2008, 49, 3588–3591. [Google Scholar]
  11. Tsutsumi, Y.; Yamakawa, K.; Yoshida, M.; Ema, T.; Sakai, T. Bifunctional organocatalyst for activation of carbon dioxide and epoxide to produce cyclic carbonate: Betaine as a new catalytic motif. Org. Lett. 2010, 12, 5728–5731. [Google Scholar]
  12. Takahashi, T.; Watahiki, T.; Kitazume, S.; Yasuda, H.; Sakakura, T. Synergistic hybrid catalyst for cyclic carbonate synthesis: Remarkable acceleration caused by immobilization of homogeneous catalyst on silica. Chem. Commun. 2006, 1664–1666. [Google Scholar]
  13. Zhou, H.; Wang, Y.M.; Zhang, W.Z.; Qu, J.P.; Lu, X.B. N-Heterocyclic carbene functionalized MCM-41 as an efficient catalyst for chemical fixation of carbon dioxide. Green Chem. 2011, 13, 644–650. [Google Scholar]
  14. Cheng, W.G.; Chen, X.; Sun, J.; Wang, J.Q.; Zhang, S.J. SBA-15 supported triazolium-based ionic liquids as highly efficient and recyclable catalysts for fixation of CO2 with epoxides. Catal. Today 2013, 200, 117–124. [Google Scholar]
  15. Wang, Y.M.; Wu, Z.Y.; Wei, Y.L.; Zhu, J.H. In situ coating metal oxide on SBA-15 in one-pot synthesis. Micropor. Mesopor. Mater. 2005, 84, 127–136. [Google Scholar]
  16. Dai, W.L.; Jin, B.; Luo, S.L.; Luo, X.B.; Tu, X.M.; Au, C.T. Novel functionalized guanidinium ionic liquids: Efficient acid-base bifunctional catalysts for CO2 fixation with epoxides. J. Mol. Catal. A 2013, 378, 326–332. [Google Scholar]
  17. Rojas, M.F.; Bernard, F.L.; Aquino, A.; Borges, J.; Vecchia, F.D.; Menezes, S.; Ligabue, R.; Einloft, S. Poly(ionic liquid)s as efficient catalyst in transformation of CO2 to cyclic carbonate. J. Mol. Catal. A 2014, 392, 83–88. [Google Scholar]
  18. Sun, J.M.; Fujita, S.I.; Zhao, F.Y.; Arai, M. A highly efficient catalyst system of ZnBr2/n-Bu4NI for the synthesis of styrene carbonate from styrene oxide and supercritical carbon dioxide. Appl. Catal. A 2005, 287, 221–226. [Google Scholar]
  19. Liu, M.S.; Liu, B.; Shi, L.; Wang, F.X.; Liang, L.; Sun, J.M. Melamine-ZnI2 as heterogeneous catalysts for efficient chemical fixation of carbon dioxide to cyclic carbonate. RSC Adv. 2015, 5, 960–966. [Google Scholar]
  20. Kim, M.I.; Choi, S.J.; Kim, D.W.; Park, D.W. Catalytic performance of zinc containing ionic liquids immobilized on silica for the synthesis of cyclic carbonates. J. Ind. Eng. Chem. 2014, 20, 3102–3107. [Google Scholar]
  21. Lee, J.K.; Kim, Y.J.; Choi, Y.S.; Lee, H.; Lee, J.S.; Hong, J.; Jeong, E.K.; Kim, H.S.; Cheong, M. Zn-containing ionic liquids bearing dialkylphosphate ligands for the coupling reactions of epoxides and CO2. Appl. Catal. B 2012, 111–112, 621–627. [Google Scholar]
  22. Wang, J.Q.; Sun, J.; Cheng, W.G.; Dong, K.; Zhang, X.P.; Zhang, S.J. Experimental and theoretical studies on hydrogen bond-promoted fixation of carbon dioxide and epoxides in cyclic carbonates. Phys. Chem. Chem. Phys. 2012, 14, 11021–11026. [Google Scholar]
  23. Xiao, L.F.; Lv, D.W.; Wu, W. Brønsted acidic ionic liquids mediated metallic salts catalytic system for the chemical fixation of carbon dioxide to form cyclic carbonates. Catal. Lett. 2011, 141, 1838–1844. [Google Scholar]
  24. Motokura, K.; Itagaki, S.; Iwasawa, Y.; Miyaji, A.; Baba, T. Zinc-accelerated cycloaddition of carbon dioxide to styrene oxide catalyzed by pyrrolidinopyridinium iodides. Top Catal. 2014, 57, 953–959. [Google Scholar]
  25. Ema, T.; Miyazaki, Y.; Shimonishi, J.; Maeda, C.; Hasegawa, J.Y. Bifunctional porphyrin catalysts for the synthesis of cyclic carbonates from epoxides and CO2: Structural optimization and mechanistic study. J. Am. Chem. Soc. 2014, 136, 15270–15279. [Google Scholar]
  26. Adam, F.; Batagarawa, M.S. Tetramethylguanidine-silica nanoparticles as an efficient and reusable catalyst for the synthesis of cyclic propylene carbonate from carbon dioxide and propylene oxide. Appl. Catal. A 2013, 454, 164–171. [Google Scholar]
  27. Miao, C.X.; Wang, J.Q.; Wu, Y.; Du, Y.; He, L.N. Bifunctional metal-salen complexes as efficient catalysts for the fixation of CO2 with epoxides under solvent-free conditions. ChemSusChem. 2008, 1, 236–241. [Google Scholar]
  28. Yang, Z.Z.; Jiang, D.E.; Zhu, X.; Tian, C.C.; Brown, S.; Thanh, C.L.; He, L.N.; Dai, S. Coordination effect-regulated CO2 capture with an alkali metal onium salts/crown ether system. Green Chem. 2014, 16, 253–258. [Google Scholar]
  29. Yue, C.T.; Su, D.; Zhang, X.; Wu, W.; Xiao, L.F. Amino-functional imidazolium ionic liquids for CO2 activation and conversion to form cyclic carbonate. Catal. Lett. 2014, 144, 1313–1321. [Google Scholar]
  30. Zhong, S.F.; Liang, L.; Liu, B.; Sun, J.M. ZnBr2/DMF as simplest highly active Acid-Base homogeneous catalysts for cycloaddition of carbon dioxide to propylene oxide. J. CO2 Util. 2014, 6, 75–79. [Google Scholar]
  31. Prasetyanto, E.A.; Ansari, M.B.; Min, B.H.; Park, S.E. Melamine tri-silsesquioxane bridged periodic mesoporous organosilica as an efficient metal-free catalyst for CO2 activation. Catal. Today 2010, 158, 252–257. [Google Scholar]

Share and Cite

MDPI and ACS Style

Liu, B.; Liu, M.; Liang, L.; Sun, J. Guanidine Hydrochloride/ZnI2 as Heterogeneous Catalyst for Conversion of CO2 and Epoxides to Cyclic Carbonates under Mild Conditions. Catalysts 2015, 5, 119-130. https://doi.org/10.3390/catal5010119

AMA Style

Liu B, Liu M, Liang L, Sun J. Guanidine Hydrochloride/ZnI2 as Heterogeneous Catalyst for Conversion of CO2 and Epoxides to Cyclic Carbonates under Mild Conditions. Catalysts. 2015; 5(1):119-130. https://doi.org/10.3390/catal5010119

Chicago/Turabian Style

Liu, Bo, Mengshuai Liu, Lin Liang, and Jianmin Sun. 2015. "Guanidine Hydrochloride/ZnI2 as Heterogeneous Catalyst for Conversion of CO2 and Epoxides to Cyclic Carbonates under Mild Conditions" Catalysts 5, no. 1: 119-130. https://doi.org/10.3390/catal5010119

Article Metrics

Back to TopTop