Next Article in Journal
The Study of Localized Crack-Induced Effects of Nonlinear Vibro-Acoustic Modulation
Next Article in Special Issue
Efficient Synthesis of 2D Mica Nanosheets by Solvothermal and Microwave-Assisted Techniques for CO2 Capture Applications
Previous Article in Journal
Multi-Energy and Fast-Convergence Iterative Reconstruction Algorithm for Organic Material Identification Using X-ray Computed Tomography
Previous Article in Special Issue
Optimization of CO2 Sorption onto Spent Shale with Diethylenetriamine (DETA) and Ethylenediamine (EDA)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solvent-Free Coupling Reaction of Carbon Dioxide and Epoxides Catalyzed by Quaternary Ammonium Functionalized Schiff Base Metal Complexes under Mild Conditions

1
Key Laboratory of Catalysis and Energy Materials Chemistry of Education & Hubei Key Laboratory of Catalysis and Materials Science, College of Chemistry and Materials Science, South-Central Minzu University, Wuhan 430074, China
2
School of Chemical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(4), 1646; https://doi.org/10.3390/ma16041646
Submission received: 16 December 2022 / Revised: 12 February 2023 / Accepted: 13 February 2023 / Published: 16 February 2023

Abstract

:
A series of bifunctional Schiff base metal catalysts (Zn-NPClR, Zn-NPXH, and M-NPClH) with two quaternary ammonium groups were prepared for carbon dioxide (CO2) and epoxide coupling reactions. The effects of the reaction variables on the catalytic activity were systematically investigated, and the optimal reaction conditions (120 °C, 1 MPa CO2, 3 h) were screened. The performances of different metal-centered catalysts were evaluated, and Co-NPClH showed excellent activity. This kind of bifunctional catalyst has a wide range of substrate applicability, excellent stability, and can be reused for more than five runs. A relatively high TOF could reach up to 1416 h−1 with Zn-NPClH as catalyst by adjusting reaction factors. In addition, the kinetic study of the coupling reaction catalyzed by three catalysts (Zn, Co, and Ni) was carried out to obtain the activation energy (Ea) for the formation of cyclic carbonates. Finally, a possible mechanism for this cyclization reaction was proposed.

1. Introduction

Since the industrial revolution, large amounts of CO2 emissions are causing its concentration in the air to increase every year; it is now about 400 ppm (mg/L), leading to a series of environmental and social problems [1,2]. However, CO2 can be used as a non-toxic, cheap, easily available, and abundant C1 resource for constructing C-C, C-N, and C-O bonds to produce high-value chemicals such as carboxylic acids, oxazolidinones, urea, isocyanates, cyclic carbonates, etc. [3,4,5,6,7,8,9]. The synthesis of cyclic carbonates using carbon dioxide has been extensively studied. One is due to the 100% atomic utilization of the reaction, and the other is that the products have favorable physicochemical properties such as environmental friendliness, high boiling point, low odor, and low toxicity, and can be used as an alternative solvent to the high boiling point polar solvents DMF and DMSO [6]. In addition, cyclic carbonate can be applied to fuel additives, electrolytes for lithium-ion batteries [10,11], and polycyclic carbonate synthesis [12,13].
The thermodynamic and kinetic inertness of CO2 make its conversion require a large energy input (e.g., high temperature and pressure), and the search for suitable catalysts is a focus of research on whether CO2 can be used efficiently [6]. From a mechanistic point of view, in the synthesis of cyclic carbonates using CO2 and epoxides, catalysts have an important role in the ring opening of epoxides: (1) acting on the oxygen atom of the ring via Lewis acid or the formation of hydrogen bonds, and (2) nucleophilic reagents attacking the carbon atom of the epoxide to promote its ring opening [14]. Based on these presumptions, many catalytic systems have been developed for applications in the cycloaddition reactions of epoxides with carbon dioxide, such as ionic liquids [15,16,17,18], N-heterocyclic carbenes (NHCs) [19], porphyrins [20], organometallic framework materials (MOF) [21], inorganic catalysts [22], metal complex catalysts [23,24], carboxylic acid compounds [25,26,27], and quaternary ammonium salt-based catalysts [28,29].
Based on the need for catalytic systems to have both electrophilic and nucleophilic centers, various “bifunctional catalysts” have been developed [30,31,32]. In this paper, a variety of quaternary ammonium salt-based bifunctional metal catalysts were designed and synthesized for the cycloaddition reaction of epoxides with carbon dioxide (Scheme 1). The results show that such bifunctional catalysts can effectively promote the formation of cyclic carbonate from epoxide and CO2 at low catalyst loadings (1 mol%) and relatively mild conditions (atmospheric pressure). In addition, since the reaction is heterogeneously catalyzed, catalyst recycling is simple and convenient. The catalyst recycling experiments and IR analysis of the catalysts showed that the prepared bifunctional catalysts can be effectively recycled five times without deactivation.

2. Materials and Methods

2.1. Chemicals and Analytical Methods

The information on materials is listed in Table S1: Provenance and mass fraction purity of the materials. All the chemicals were used as received except for the epoxides, which were purified by distillation from CaH2 before utilization. The general procedure for the cycloaddition of CO2 to epoxides and spectra copies of all synthesized complexes and carbonate products are provided in the Supporting Information (SI). The yield and selectivity are determined by 1H NMR characterization.

2.2. Synthesis of Quaternary Ammonium Modified Metal Complexes Zn-NPClR, Zn-NPXH, M-NPClH

The synthetic route for catalysts is shown in Scheme 2.

2.2.1. Synthesis of 5-Chloromethyl Salicylaldehyde (PX, NPX)

The ligand precursors PX and NPX (X = Cl, Br, I) were prepared and characterized according to the method reported by Ji et al. (Scheme 2) [33,34]. For comparison, a complex C0 was synthesized as a non-bifunctional catalyst.

2.2.2. General Procedure for Synthesis of Quaternary Ammonium Modified Metal Complexes Zn-NPClR, Zn-NPXH

To a 100 mL dry round-bottom flask equipped with a magnetic stir bar, an ethanol solution (20 mL) of quaternary ammonium modified salicylaldehyde NPX (0.01186 mol) and an ethanol solution (10 mL) of 0.01533 mol amine C6H6R (R = H, Cl, NO3, CH3, C4H9) were added sequentially. The obtained mixture was heated and refluxed for 4 h. Then zinc acetate anhydrous (7.74 mmol) was added to the mixture. After another 6 h reflux, a large amount of precipitates were obtained. These precipitates were filtered and washed with ethanol several times, then dried in vacuum to obtain quaternary ammonium-modified salicylaldehyde-zinc catalysts. Details of these complexes are shown in SI.

2.2.3. Synthesis of Other Metal Center (Cobalt, Plumbum, Nickel, Copper) Complexes M-NPClH

Following the procedure of Zn-NPClH, we have also prepared different metal-centered quaternary ammonium salt complexes using other metal salts (cobalt chloride hexahydrate, lead acetate, nickel chloride hexahydrate, copper chloride dihydrate). The physicochemical information of these complexes is shown in Supporting Information.

2.3. Catalytic Performance for CO2/Epoxide Cycloaddition Reaction

The cycloaddition reaction of CO2 and epoxide was conducted in a 50 mL stainless steel autoclave. The required catalyst and epoxide were added, and then CO2 was pressurized into the reactor at a certain pressure. The autoclave was immersed in an oil bath at a preset temperature, with stirring. After a proper time, the autoclave was cooled down to room temperature and then vented slowly. The resulting mixture was analyzed by 1H NMR to give the yield and the selectivity.

3. Results and Discussion

3.1. The Effect of Time on Cycloaddition of Propylene Oxide (PO) and CO2

To optimize the reaction conditions for the cyclization reaction of epoxide and carbon dioxide, Zn-NPClH was used as a catalyst and PO as a substrate. The effect of time on the reaction was investigated at 120 °C and 1 MPa CO2, as shown in Figure 1A. The propylene carbonate (PC) content increased steeply with time, reaching 3 h. It is obvious that within the initial 3 h, both the reaction rate and the PC yield increased rapidly. After 3 h, the PC yield still increased, but not as dramatically as before. The possible reason is that as the reaction proceeds, the PC concentration increases and the PO concentration decreases; in addition, the generated PC increases the viscosity of the system, which leads to a reduced opportunity for contact between the catalytically active site and the reaction substrate [35]. Therefore, 3 h was chosen as the optimal reaction time.

3.2. The Effect of Temperature on Cycloaddition of PO and CO2

The PC yield was low at 100 °C (Figure 1B), while increasing the temperature greatly improved the catalytic activity, which confirms the fact that the temperature has a significant effect on the reaction activity, and the fact that a high temperature can promote the PO conversion also indicates that this cycloaddition reaction is thermodynamically favorable [36,37,38]. When the temperature was increased from 100 °C to 120 °C, the PC yield increased steeply from 16% to 93%, but there was no significant change in the PC yield when the temperature was continued to increase to 140 °C (Figure 1B). This is because the cycloaddition reaction of CO2 and PO is exothermic, so too high a temperature will hinder the formation of cyclic carbonate. In addition, high temperatures will also lead to the polymerization of cyclic carbonate [39,40]. Based on the above study, 120 °C was chosen as the optimal reaction temperature.

3.3. The Effect of CO2 Pressure on Cycloaddition of PO and CO2

CO2 pressure affects the mass transfer kinetics of this cycloaddition reaction as well as the yield and selectivity of PC [41], so the relationship between pressure and yield was explored in the range of 0.5–5 MPa (Figure 1C). The yield of PC increased from 35% to 96%when the pressure was increased from 0.5 MPa to 1 MPa and decreased slightly to 94% when the pressure was continually increased to 2 MPa. Further raising the reaction pressure to 3 MPa and 5 MPa, the PC yield lowered to 73% and 66%, respectively. The increase in pressure makes the concentration of CO2 in PO higher, which promotes the reaction. However, too high pressure will reduce the concentration of PO in the vicinity of the catalyst, which will bring down the catalytic activity and thus reduce the reaction rate and the PC yield [42,43]. Experimental results show that the optimal CO2 pressure for this reaction is 1 MPa.

3.4. The Effect of the Ratio of PO to Catalyst on Cycloaddition of PO and CO2

Figure 1D shows the effect of catalyst loadings on PC yield at 120 °C and 1 MPa CO2. When the amount of catalyst was varied from 0.1 to 1 mol%, the PC yield dramatically improved from 10.2% to 93.2%. As the catalyst dosage increases, more catalytic active centers can be supplied to the catalytic system, so the PC yield is significantly increased [44]. However, further increasing the amount of the catalyst cannot lead to a big increase in the yield of carbonate; this may be due to the fact that the presence of excess catalyst makes its dispersion low in the system, thus hindering mass transfer between the catalyst active center and the reactants [45,46]. Therefore, the optimal ratio of catalyst to PO is 1:100.

3.5. Cycloaddition of CO2 to PO by Various Bifunctional Catalysts

The catalytic performances of various bifunctional catalysts were investigated for the cycloaddition reaction of PO and CO2 under optimal conditions (120 °C, 1 MPa, 3 h, 1:100). The corresponding results were summarized in Table 1. With the non-bifunctional complex C0 or Et3N as the sole catalyst, very low activity was observed (Table 1, entries 1,2). Using TBAC/Et3N or C0/Et3N as two-component catalysts, the catalytic activity was clearly improved (Table 1, entries 4,5). However, under the same conditions, a high yield of 93.2% was obtained using a quaternary ammonium salt bifunctional complex of Zn-NPClH as the catalyst (Table 1, entry 6). The electron effect of different substituents on ligands can affect the catalytic activity, so the effect of different ligands with electron-donating or electron-withdrawing substituents on the catalytic results was investigated. Under the same condition, the catalytic activity of the catalyst differs as follows: Zn-NPClNO2 > Zn-NPClCl > Zn-NPClH > Zn-NPClCH3 > Zn-NPClC4H9 (-NO2 > -Cl > -H > -CH3 > -C4H9) (Table 1, entries 6–10). This result demonstrates that the catalytic effect was related to the electronegativity of the substituents on the catalyst, and catalysts with a strong electron-withdrawing substituent showed better catalytic activity. It could be attributed to the ability of the strong withdrawing substituent, which can reduce the electron cloud density of the metal complexes and enhance their Lewis acidity. We also studied the effect of different metal centers in coordination compounds on catalytic activity (Table 1, entries 6,11–14). The results showed that cobalt, zinc, lead, and nickel metal center complexes have relatively good catalytic activity but poor activity for the copper center; the activity order is Co-NPClH > Zn-NPClH > Pb-NPClH > Ni-NPClH > Cu-NPClH (Co > Zn > Pb > Ni > Cu). The catalytic activity of different metal centers is highly dependent on the metal center type, which may be related to the different coordination abilities of metal centers with PO, while the low activity of bifunctional complexes with Cu metal centers may be attributed to the low acidity of Cu centers [47,48]. The anion plays an important role in the reaction of the epoxide with CO2 by nucleophilically attacking the epoxide to form a C-X bond to open its ring, so the anion species are screened. Accordingly, the more easily departed anion exhibits the best catalytic effect, which is consistent with the experimental results: I > Br > Cl (Table 1, entries6, 15, 16). To our delight, by adjusting reaction factors (0.1 mol% Zn-NPClH, 150 °C, 0.5 h), a relatively high TOF could reach up to 1416 h1 with Zn-NPClH as catalyst (Table 1, entry 17).

3.6. Study of Catalytic Performance at Atmospheric Pressure

The catalytic activity of these bifunctional catalysts at atmospheric pressure has also been studied to meet the needs of sustainable development and green industrialization. The cycloaddition reaction of CO2 with isopropyl glycidyl ether was used as a model reaction to study the catalytic activity of the bifunctional metal catalysts M-NPClH at 0.1 MPa (Figure 2). All catalysts showed good to excellent activity. Overall, the order of catalytic activity of different metals at atmospheric pressure was consistent with that at high pressure: Co-NPClH > Zn-NPClH > Pb-NPClH > Ni-NPClH > Cu-NPClH. Zn-NPClH, Co-NPClH, Pb-NPClH, and Ni-NPClH all catalyzed glycidyl isopropyl ether better, with the catalyst Co-NPClH reaching 96.4% yield after 14 h of continuous reaction.

3.7. Substrate Scope of CO2/Epoxide Coupling

In order to investigate the generalizability of the bifunctional catalyst to other cycloaddition reactions of epoxides with CO2, we examined the adaptation of the bifunctional catalyst Zn-NPClH to a variety of epoxyalkane substrates under the above-mentioned optimized conditions, and the results are shown in Table 2 (Condition A). The results show that the catalyst system is effective in converting the studied common epoxides to their corresponding cyclic carbonates. For terminal epoxides, better yields of substrates with either electron-absorbing or electron-donating groups attached were obtained with the activation of the catalyst (Table 2, entries 1–2, 7–9). The nucleophilic attack of the epoxide by chloride ions is hindered by the high spatial site resistance of the epoxide environment, resulting in very low yields of dimethyloxirane, cyclohexene oxide, and 1,2-epoxyphenylethane (Table 2, entries 3–6), as previously reported in the literature [49,50,51]. In addition, the epoxide substrates of the glycidyl ether series can also be coupled with carbon dioxide, as in the case of dicyclic oxides, which can also be well formed as cyclic carbonates of the corresponding dicyclic ring. It is worth pointing out that bicyclic carbonates can be used as raw materials to produce non-isocyanate polyurethanes (NIPUs) by reacting with polyfunctional primary amines, which have important applications in industry [52,53].
The application of various epoxides with a higher boiling point (>110 °C) to couple with CO2 was also tested at atmospheric pressure. Although a longer time was necessary for the challenging epoxides (Table 2, Condition B), these bifunctional catalysts are generally still workable for a wide scope of epoxides, even under 0.1 MPa.

3.8. Catalyst Recycling

The reuse of catalysts in industry will not only reduce costs, but their reusability also reflects the stability of the catalyst under reaction conditions. Using Zn-PPBCl as a catalyst to catalyze the formation of PC from CO2 and PO as a model reaction, its reusability performance under optimal conditions was investigated. The synthesized Zn-PPBCl could be reused at least five times without significant loss of catalytic activity and selectivity (Figure 3). In addition, to further investigate the stability of this catalyst, the structures of both the5-cycleand fresh catalysts were also characterized by IR. As shown in Figure 4, the structure of the catalyst was maintained after five times of reuse, which proves that the synthesized catalyst is not only efficient but also stable and reusable.

3.9. Kinetic Studies

Kinetics study of the cycloaddition reaction using glycidyl isopropyl ether (GIE) as substrate using three bifunctional quaternary ammonium metal catalysts M-NPClH (M = Zn, Co, Ni) at 1 atm CO2 pressure (see the ESI, Table S2: Structure of ligands–S15: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Ni-NPClH at 393 K). The activation energies (Ea) for the three catalysts were 30.8 kJ/mol (Zn), 26.1 kJ/mol (Co), and 32.9 kJ/mol (Ni), respectively, which were consistent with the order of catalytic activity (Figure 4, Figure 5 and Figure 6). The higher conversion obtained with Co-NPClH could likely be attributed to the higher Lewis acidity of the Co center compared with other metal centers [54].

4. Proposed Reaction Mechanism

Based on the experimental results, kinetic study analysis, and previously reported literature [53,54], we proposed a possible mechanism for the cycloaddition reaction of CO2 with epoxide (Scheme 3). The metal center of the catalyst activates the oxygen on the epoxide, and the halogen ion on the functional group of the bifunctional metal complex catalyst nucleophilically attacks the less site-resistant carbon on the epoxide, causing the epoxide to undergo a ring-opening step to form a metal alcohol salt intermediate. At the same time, CO2 is inserted into this intermediate to form a metal carboxylate. Finally, the oxygen anion of the intermediate attacks the C-X bond, closing the loop to form propylene carbonate while completing the catalytic cycle to release the catalyst.

5. Conclusions

The coupling reactions of various epoxides with CO2 were successfully achieved by using halogenated quaternary ammonium salt metal complex catalysts without the addition of co-catalysts and solvents. When using Co-NPClH as a catalyst to promote the reaction with PO, the conversion was 98.6% and the PC selectivity was 99% under the optimal conditions (120 °C, 1 MPa CO2, and 1.0 mol% Co-NPClH). In addition, the synthesized bifunctional catalysts can be reused by simple operations and reused at least five times with essentially no decrease in activity or selectivity. The kinetic studies using catalysts M-NPClH (M = Co, Zn, Ni) are documented for the cycloaddition of GIE and CO2, in which the apparent activation energies (Ea) are 26.1, 30.8, and 32.9 kJ/mol, respectively. The series of bifunctional metal catalysts synthesized in this study exhibited excellent catalytic performance, high TOF values, and strong stability, which proved the potential of these catalysts for industrial applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16041646/s1, Table S1: Provenance and mass fraction purity of the materials. Table S2: Structure of ligands. Table S3: Structure of catalysts. Figure S1: 1H NMR and 13C NMR spectra for NPCl in DMSO-d6. Figure S2: 1H NMR and 13C NMR spectra for Zn-NPClH in DMSO-d6. Figure S3: 1H NMR spectrum for Zn-NPClCH3 in DMSO-d6. Figure S4: 1H NMR and 13C NMR spectra for Zn-NPClCl in DMSO-d6. Figure S5: 1H NMR spectrum for Zn-NPClNO2 in DMSO-d6. Figure S6: 1H NMR and 13C NMR spectra for Zn-NPClC4H9 in DMSO-d6. Table S4. Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Zn-NPClH at 363 K. Table S5: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Zn-NPClH at 373 K. Table S6: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Zn-NPClH at 383 K. Table S7: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Zn-NPClH at 393 K. Figure S7: Logarithmic plots of (1 − x) versus time by Zn-NPClH at 363–393 K. Table S8: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Co-NPClH at 363 K. Table S9: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Co-NPClH at 373 K. Table S10: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Co-NPClH at 383 K. Table S11: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Co-NPClH at 393 K. Figure S8: Logarithmic plots of (1 − x) versus time by Co-NPClH at 363–393 K. Table S12: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Ni-NPClH at 363 K. Table S13: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Ni-NPClH at 373 K. Table S14: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Ni-NPClH at 383 K. Table S15: Effects of the reaction time on cycloaddition of CO2 and glycidyl isopropyl ether at atmospheric pressure by Ni-NPClH at 393 K. Figure S9. Logarithmic plots of (1 − x) versus time by Ni-NPClH at 363–393 K.

Author Contributions

Q.W.: validation, formal analysis, resources, writing—original draft preparation. X.Y.: methodology and data curation. Q.Z.: data curation and methodology. H.-J.Y.: conceptualization, validation, writing—review and editing, project administration, funding acquisition. Q.J.: software, visualization. J.H.: investigation and visualization. C.-Y.G.: conceptualization, validation, supervision, project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 51073175.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data presented in this study are openly available in MDPI.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pera-Titus, M. Porous inorganic membranes for CO2 capture: Present and prospects. Chem. Rev. 2014, 114, 1413–1492. [Google Scholar] [CrossRef] [PubMed]
  2. Qu, R.; Ren, Z.-j.; Li, N.; Zhang, F.-F.; Zhang, Z.-J.; Zhang, H.-N. Solvent-free cycloaddition of carbon dioxide and epichlorohydrin catalyzed by surface-attached imidazolium-type poly(ionic liquid) monolayers. J. CO2 Util. 2020, 38, 168–176. [Google Scholar] [CrossRef]
  3. Huang, H.; Ye, J.-H.; Zhu, L.; Ran, C.-K.; Miao, M.; Wang, W.; Chen, H.-J.; Zhou, W.-J.; Lan, Y.; Yu, B.; et al. Visible-light-driven anti-markovnikov hydrocarboxylation of acrylates and styrenes with CO2. CCS Chem. 2020, 2, 1746–1756. [Google Scholar]
  4. Kumar, A.; Sharma, P.; Sharma, N.; Kumar, Y.; Mahajan, D. Catalyst free N-formylation of aromatic and aliphatic amines exploiting reductive formylation of CO2 using NaBH4. RSC Adv. 2021, 11, 25777–25787. [Google Scholar] [CrossRef]
  5. Gong, Y.-X.; Li, Y.-F.; Hu, J.-Y.; Wang, Z.-G.; Deng, T.-L. Sulfur-containing amino acid-derived ionic liquid as a halogen-free catalyst for CO2 mild transformation into cyclic carbonates. New J. Chem. 2021, 45, 19215–19218. [Google Scholar] [CrossRef]
  6. Ge, Y.-S.; Liu, W.; Zou, Y.-Z.; Cheng, G.; Ke, H.-Z. A solid Zn complex catalyst for efficient transformation of CO2 to cyclic carbonates at mild conditions. Tetrahedron 2022, 119, 132857–132864. [Google Scholar] [CrossRef]
  7. Roy, P.; Pramanik, A.; Sarkar, P. Dual-silicon-doped graphitic carbon nitride sheet: An efficient metal-free electrocatalyst for urea synthesis. J. Phys. Chem. Lett. 2021, 12, 10837–10844. [Google Scholar] [CrossRef]
  8. Zhang, Z.; Zhang, Z.-H.; Zhou, F.; Zhou, J. Catalytic enantioselective transfer hydrogenation-carboxylative cyclization to 4-fluoroalkyl 2-oxazolidinone with CO2 as the C1 synthon. Org. Lett. 2021, 23, 2726–2730. [Google Scholar] [CrossRef]
  9. Gu, A.-L.; Wang, W.-T.; Cheng, X.-Y.; Hu, T.-D.; Wu, Z.-L. Non-noble-metal metal-organic-framework-catalyzed carboxylative cyclization of propargylic amines with atmospheric carbon dioxide under ambient conditions. Inorg. Chem. 2021, 60, 13425–13433. [Google Scholar] [CrossRef]
  10. Xu, K. Nonaqueous liquid electrolytes for lithium-based rechargeable batteries. Chem. Rev. 2004, 104, 4303–4417. [Google Scholar] [CrossRef]
  11. Zhao, C.; Luo, X.; Chen, C.; Wu, H. Sandwich electrode designed for high performance lithium-ion battery. Nanoscale 2016, 8, 9511–9516. [Google Scholar] [CrossRef] [PubMed]
  12. Matsukizono, H.; Endo, T. Reworkable polyhydroxy urethane films with reversible acetal networks obtained from multifunctional six-membered cyclic carbonates. J. Am. Chem. Soc. 2018, 140, 884–887. [Google Scholar] [CrossRef] [PubMed]
  13. Motokucho, S.; Takenouchi, Y.; Satoh, R.; Morikawa, H.; Nakatani, H. Novel polyurethane-catalyzed cyclic carbonate synthesis using CO2 and epoxide. ACS Sustain. Chem. Eng. 2020, 8, 4337–4340. [Google Scholar] [CrossRef]
  14. Peng, J.; Yang, H.J.; Wei, Z.D.; Guo, C.-Y. Metal complexes bearing 2-(imidazol-2-yl)phenol ligands: Synthesis, characterization and catalytic performance in the fixation of carbon dioxide with epoxides. RSC Adv. 2015, 5, 53063–53072. [Google Scholar] [CrossRef]
  15. Liu, M.; Liang, L.; Li, X.; Gao, X.; Sun, J. Novel urea derivative-based ionic liquids with dual-functions: CO2 capture and conversion under metal- and solvent-free conditions. Green Chem. 2016, 18, 2851–2863. [Google Scholar] [CrossRef]
  16. Peng, Y.; Song, N.; Heng, N.; Yang, H. Coupling reaction of CO2 and epoxides catalyzed byimidazole functionalized iron complexes. J. South-Central Minzu Univ. (Nat. Sci. Ed.) 2021, 1, 9–14. [Google Scholar]
  17. Xie, Y.; Liang, J.; Fu, Y.; Huang, M.; Xu, X.; Wang, H.; Tu, S.; Li, J. Hypercrosslinked mesoporous poly(ionic liquid)s with high ionic density for efficient CO2 capture and conversion into cyclic carbonates. J. Mater. Chem. A 2018, 6, 6660–6666. [Google Scholar] [CrossRef]
  18. Zhang, Y.-W.; El-Sayed, E.M.; Su, K.-Z.; Yuan, D.-Q.; Han, Z.-B. Facile syntheses of ionic polymers for efficient catalytic conversion of CO2 to cyclic carbonates. J. CO2 Util. 2020, 42, 101301–101306. [Google Scholar] [CrossRef]
  19. Peng, J.; Yang, H.J.; Geng, Y.C.; Wei, Z.D.; Wang, L.H.; Guo, C.Y. Novel, recyclable supramolecular metal complexes for the synthesis of cyclic carbonates from epoxides and CO2 under solvent-free conditions. J. CO2 Util. 2017, 17, 243–255. [Google Scholar] [CrossRef]
  20. Maeda, C.; Shimonishi, J.; Miyazaki, R.; Hasegawa, J.Y.; Ema, T. Highly active and robust metalloporphyrin catalysts for the synthesis of cyclic carbonates from abroad range of epoxides and carbon dioxide. Chem. Eur. J. 2016, 22, 6556–6563. [Google Scholar] [CrossRef]
  21. Han, Y.-H.; Zhou, Z.-Y.; Tian, C.-B.; Du, S.-W. A dual-walled cage MOF as an efficient heterogeneous catalyst for the conversion of CO2 under mild and co-catalyst free conditions. Green Chem. 2016, 18, 4086–4091. [Google Scholar] [CrossRef] [Green Version]
  22. Peng, J.; Yang, H.-J.; Wang, S.; Ban, B.; Wei, Z.; Lei, B.; Guo, C.-Y. Efficient solvent-free fixation of CO2 catalyzed by new recyclable bifunctional metal complexes. J. CO2 Util. 2018, 24, 1–9. [Google Scholar] [CrossRef]
  23. 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] [CrossRef] [PubMed]
  24. Li, P.; Cao, Z. Catalytic preparation of cyclic carbonates from CO2 and epoxides by metaleporphyrin and corrole complexes: Insight into effects of cocatalyst and meso-substitution. Organometallics 2018, 37, 406–414. [Google Scholar] [CrossRef]
  25. Yang, H.; Nie, H.; Wen, Q.; Ban, B.; Hu, J. Cycloaddition reaction of azridinane and CO2 catalyzed by bifunctional zinc catalyst. J. Xuzhou Inst. Technol. (Nat. Sci. Ed.) 2022, 37, 28–35. [Google Scholar]
  26. Wang, Y.-J.; Shen, Y.-H.; Wang, Z. Rapid conversion of CO2 and propylene oxide into propylene carbonate over acetic acid/KI under relatively mild conditions. New J. Chem. 2021, 45, 20323–20328. [Google Scholar] [CrossRef]
  27. Rostami, A.; Ebrahimi, A.; Al-Jassasi, M.; Mirzaei, S.; Al-Harrasi, A. 2-Picolinic acid as a naturally occurring hydrogen bond donor for the preparation of cyclic carbonates from terminal/internal epoxides and CO2. Green Chem. 2022, 24, 9069–9083. [Google Scholar] [CrossRef]
  28. Cheng, W.-G.; Xiao, B.-N.; Zhang, S.-J. Effect of hydrogen bond of hydroxyl-functionalized ammonium ionic liquids on cycloaddition of CO2. Tetrahedron Lett. 2015, 56, 1416–1419. [Google Scholar] [CrossRef]
  29. Hong, M.; Kim, Y.; Kim, H.; Cho, H.J.; Baik, M.H.; Kim, Y. Scorpionate catalysts for coupling CO2 and epoxides to cyclic carbonates: A rational design approach for organocatalysts. J. Org. Chem. 2018, 83, 9370–9380. [Google Scholar] [CrossRef]
  30. Luo, R.-C.; Yang, Z.; Zhang, W.-Y.; Zhou, X.-T.; Ji, H.-B. Recyclable bifunctional aluminum salen catalyst for CO2 fixation: The efficient formation of five-membered heterocyclic compounds. Sci. China Chem. 2017, 60, 979–989. [Google Scholar] [CrossRef]
  31. Liu, L.-N.; Jayakumar, S.; Chen, J.; Tao, L.; Li, H.; Yang, Q.-H.; Li, C. Synthesis of bifunctional porphyrin polymers for catalytic conversion of dilute CO2 to cyclic carbonates. ACS Appl. Mater. Interfaces 2021, 13, 29522–29531. [Google Scholar] [CrossRef] [PubMed]
  32. Mitra, A.; Ghosh, S.; Paliwal, K.S.; Ghosh, S.; Tudu, G.; Chandrasekar, A.; Mahalingam, V. Alumina-based bifunctional catalyst for efficient CO2 fixation into epoxides at atmospheric pressure. Inorg. Chem. 2022, 61, 16356–16369. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, H.; Nie, H.; Wen, Q.; Ban, B.; Hu, J. Reaction study on CO2 with amines catalyzed by N,N-dimethylethanolamine. J. Xuzhou Inst. Technol. (Nat. Sci. Ed.) 2022, 1, 9–13. [Google Scholar]
  34. Zhang, Q.; Yu, P.; Lei, B.; Yang, H.-J.; Hu, J.; Wang, L.; Zhou, T.; Jiang, Q.; Zhu, J.; Han, B. Efficient solvent-free synthesis of cyclic carbonates from the cycloaddition of carbon dioxide and epoxides catalyzed by new imidazolinium functionalized metal complexes under 0.1 MPa. Catal. Lett. 2020, 150, 2537–2548. [Google Scholar] [CrossRef]
  35. Bai, D.-S.; Jing, H.-W.; Wang, G.-J. Cyclic carbonate synthesis from epoxides and CO2 over cyanocobalamin/n-Bu4NI. Appl. Organomet. Chem. 2012, 26, 600–603. [Google Scholar] [CrossRef]
  36. Wei, R.-J.; Zhang, X.-H.; Du, B.-Y.; Fan, Z.-Q.; Qi, G.-R. Synthesis of bis(cyclic carbonate) and propylene carbonate via a one-pot coupling reaction of CO2, bisepoxide and propylene oxide. RSC Adv. 2013, 3, 17307–17313. [Google Scholar] [CrossRef]
  37. Pescarmona, P.P.; Taherimehr, M. Challenges in the catalytic synthesis of cyclic and polymeric carbonates from epoxides and CO2. Catal. Sci. Technol. 2012, 2, 2169–2187. [Google Scholar] [CrossRef]
  38. Peng, J.; Geng, Y.; Yang, H.-J.; He, W.; Wei, Z.-D.; Yang, J.-K.; Guo, C.-Y. Efficient solvent-free fixation of CO2 into cyclic carbonates catalyzed by Bi(III) porphyrin/TBAI at atmospheric pressure. Mol. Catal. 2017, 432, 37–46. [Google Scholar] [CrossRef]
  39. Han, L.; Li, H.; Choi, S.J.; Park, M.S.; Lee, S.M.; Kim, Y.J.; Park, D.W. Ionic liquids grafted on carbon nanotubes as highly efficient heterogeneous catalysts for the synthesis of cyclic carbonates. Appl.Catal. A Gen. 2012, 429, 67–72. [Google Scholar] [CrossRef]
  40. Xu, J.; Wu, F.; Jiang, Q.; Li, Y.-X. Mesoporous carbon nitride grafted with n-bromobutane: A high-performance heterogeneous catalyst for the solvent-free cycloaddition of CO2 to propylene carbonate. Catal. Sci. Technol. 2015, 5, 447–454. [Google Scholar] [CrossRef]
  41. Son, Y.S.; Park, S.W.; Park, D.W.; Oh, K.J.; Kim, S.S. Absorption of carbon dioxide into glycidyl methacrylate solution containing the triethylamine immobilized ionic liquid on MCM-41 catalyst. Korean J.Chem. Eng. 2009, 2, 783–790. [Google Scholar] [CrossRef]
  42. Xie, Y.; Zhang, Z.-F.; Jiang, T.; He, J.-L.; Han, B.-X.; Wu, T.-B.; Ding, K.-L. CO2 cycloaddition reactions catalyzed by an ionic liquid grafted onto a highly cross-linked polymer matrix. Angew. Chem. Int. Ed. 2007, 46, 7255–7258. [Google Scholar] [CrossRef]
  43. Kilic, A.; Durgun, M.; Aytar, E.; Yavuz, R. Synthesis and characterization of novel positively charged organocobaloximes as catalysts for the fixation of CO2 to cyclic carbonates. J. Organomet. Chem. 2018, 858, 78–88. [Google Scholar] [CrossRef]
  44. Xu, J.; Xu, M.; Wu, J.; Wu, H.; Zhang, W.-H.; Li, Y.-X. Graphene oxide immobilized with ionic liquids: Facile preparation and efficient catalysis for solvent-free cycloaddition of CO2 to propylene carbonate. RSC Adv. 2015, 5, 72361–72368. [Google Scholar] [CrossRef]
  45. Sun, J.; Cheng, W.-G.; Fan, W.; Wang, Y.-H.; Meng, Z.-Y.; Zhang, S.-J. Reusable and efficient polymer-supported task-specific ionic liquid catalyst for cycloaddition of epoxide with CO2. Catal. Today 2009, 148, 361–367. [Google Scholar] [CrossRef]
  46. 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] [CrossRef]
  47. Gao, J.; Song, Q.W.; He, L.N.; Liu, C.; Yang, Z.Z.; Han, X.; Li, X.D.; Song, Q.C. Preparation of polystyrene-supported Lewis acidic Fe(III) ionic liquid and its application in catalytic conversion of carbon dioxide. Tetrahedron 2012, 68, 3835–3842. [Google Scholar] [CrossRef]
  48. Liu, M.-S.; Gao, K.-Q.; Liang, L.; Sun, J.-M.; Sheng, L.; Arai, M. Experimental and theoretical insights into binary Zn-SBA-15/KI catalysts for the selective coupling of CO2 and epoxides into cyclic carbonates under mild conditions. Catal. Sci. Technol. 2016, 6, 6406–6416. [Google Scholar] [CrossRef] [Green Version]
  49. Wang, J.-Q.; Song, W.-H.; Yi, G.-S.; Zhang, Y.-G. Imidazolium salt-modified porous hypercrosslinked polymers for synergistic CO2 capture and conversion. Chem. Commun. 2015, 51, 12076–12079. [Google Scholar] [CrossRef]
  50. Song, T.-Q.; Dong, J.; Gao, H.-L.; Cui, J.-Z.; Zhao, B. A unique zinc-organic framework constructed through in situ ligand synthesis for conversion of CO2 under mild conditions and as a luminescence sensor for Cr2O72−/CrO42−. Dalton Trans. 2017, 46, 13862–13868. [Google Scholar] [CrossRef]
  51. Chen, Y.-J.; Luo, R.-C.; Xu, Q.-H.; Jiang, J.; Zhou, X.-T.; Ji, H.-B. Metalloporphyrin polymers with intercalated ionic liquids for synergistic CO2 fixation via cyclic carbonate production. ACS Sustain. Chem. Eng. 2018, 6, 1074–1082. [Google Scholar] [CrossRef]
  52. Sheng, X.-F.; Ren, G.-J.; Qin, Y.-S.; Chen, X.-S.; Wang, X.-H.; Wang, F.-S. Quantitative synthesis of bis(cyclic carbonate)s by iron catalyst for non-isocyanate polyurethane synthesis. Green Chem. 2015, 17, 373–379. [Google Scholar] [CrossRef]
  53. Wang, T.-T.; Xie, Y.; Deng, W.-Q. Reaction mechanism of epoxide cycloaddition to CO2 catalyzed by salen-M (M = Co, Al, Zn). J. Phys. Chem. A 2014, 118, 9239–9243. [Google Scholar] [CrossRef] [PubMed]
  54. Monica, F.D.; Kleij, A.W. Mechanistic guidelines in nonreductive conversion of CO2: The case of cyclic carbonates. Catal. Sci. Technol. 2020, 10, 3483–3501. [Google Scholar] [CrossRef]
Scheme 1. Cycloaddition of CO2 to epoxides with bifunctional quaternary ammonium metal complexes as catalysts.
Scheme 1. Cycloaddition of CO2 to epoxides with bifunctional quaternary ammonium metal complexes as catalysts.
Materials 16 01646 sch001
Scheme 2. Synthesis of the quaternary ammonium catalysts.
Scheme 2. Synthesis of the quaternary ammonium catalysts.
Materials 16 01646 sch002
Figure 1. Effects of different reaction parameters on PC yield. (A) Reaction time effect on PC yield: PO (10 mL,0.143 mol), catalyst Zn-NPClH (1.43 mmol,1 mol%), T = 120 °C, P = 1 MPa; (B) Reaction temperature effect on PC yield: PO (10 mL, 0.143 mol), catalyst Zn-NPClH (1.43 mmol, 1 mol%), t = 3 h, P = 1 MPa; (C) CO2 pressure effect on PC yield: PO (10 mL, 0.143 mol), catalyst Zn-NPClH (1.43 mmol,1 mol%), t = 3 h, T = 120 °C; (D) The mole ratio of PO to catalyst effect on PC yield: PO (10 mL, 0.143 mol), catalyst Zn-NPClH (1.43 mmol), t = 3 h, T = 120 °C, P = 1 MPa. The selectivity of PC was >99%.
Figure 1. Effects of different reaction parameters on PC yield. (A) Reaction time effect on PC yield: PO (10 mL,0.143 mol), catalyst Zn-NPClH (1.43 mmol,1 mol%), T = 120 °C, P = 1 MPa; (B) Reaction temperature effect on PC yield: PO (10 mL, 0.143 mol), catalyst Zn-NPClH (1.43 mmol, 1 mol%), t = 3 h, P = 1 MPa; (C) CO2 pressure effect on PC yield: PO (10 mL, 0.143 mol), catalyst Zn-NPClH (1.43 mmol,1 mol%), t = 3 h, T = 120 °C; (D) The mole ratio of PO to catalyst effect on PC yield: PO (10 mL, 0.143 mol), catalyst Zn-NPClH (1.43 mmol), t = 3 h, T = 120 °C, P = 1 MPa. The selectivity of PC was >99%.
Materials 16 01646 g001
Figure 2. The quaternary ammonium catalyst system reaction of carbon dioxide and glycidyl isopropyl ether at atmospheric pressure. Reaction conditions: glycidyl isopropyl ether (5 mL, 0.071 mol), catalyst Zn-NPClH (0.71 mmol, 1 mol%), T = 120 °C, P = 0.1 MPa. The selectivity of PC was >99%.
Figure 2. The quaternary ammonium catalyst system reaction of carbon dioxide and glycidyl isopropyl ether at atmospheric pressure. Reaction conditions: glycidyl isopropyl ether (5 mL, 0.071 mol), catalyst Zn-NPClH (0.71 mmol, 1 mol%), T = 120 °C, P = 0.1 MPa. The selectivity of PC was >99%.
Materials 16 01646 g002
Figure 3. Reutilization of the Zn-NPClH catalyst. PO (10 mL, 0.143 mol), catalyst Zn-NPClH (1.43 mmol, 1 mol%), t = 3 h, T = 120 °C, P = 1 MPa. The selectivity of PC was >99%.
Figure 3. Reutilization of the Zn-NPClH catalyst. PO (10 mL, 0.143 mol), catalyst Zn-NPClH (1.43 mmol, 1 mol%), t = 3 h, T = 120 °C, P = 1 MPa. The selectivity of PC was >99%.
Materials 16 01646 g003
Figure 4. The FT−IR spectra of fresh catalyst and that used after five runs.
Figure 4. The FT−IR spectra of fresh catalyst and that used after five runs.
Materials 16 01646 g004
Figure 5. Arrhenius plots for the formation of cyclic carbonate from the cycloaddition of CO2 and GIE using catalysts M−NPClH.
Figure 5. Arrhenius plots for the formation of cyclic carbonate from the cycloaddition of CO2 and GIE using catalysts M−NPClH.
Materials 16 01646 g005
Figure 6. The activation energies (Ea) for the cycloaddition of CO2 and GIE catalyzed by M-NPClH.
Figure 6. The activation energies (Ea) for the cycloaddition of CO2 and GIE catalyzed by M-NPClH.
Materials 16 01646 g006
Scheme 3. Proposed mechanism for the coupling reaction of CO2 with epoxide.
Scheme 3. Proposed mechanism for the coupling reaction of CO2 with epoxide.
Materials 16 01646 sch003
Table 1. Effect of different kinds of catalyst systems on the coupling of CO2 with PO a.
Table 1. Effect of different kinds of catalyst systems on the coupling of CO2 with PO a.
EntryCatalystYield (%) bTON cTOF (h−1) dSelectivity (%) b
1C02.52.50.8>99
2Et3N0.60.60.2>99
3TBAC20.320.36.8>99
4TBAC/Et3N40.440.413.5>99
5C0/Et3N68.868.822.9>99
6Zn-NPClH93.293.231.1>99
7Zn-NPClCl95.495.431.8>99
8Zn-NPClNO296.996.932.3>99
9Zn-NPClCH390.890.830.3>99
10Zn-NPClC4H989.589.529.8>99
11Co-NPClH98.698.632.9>99
12Pb-NPClH89.089.029.7>99
13Ni-NPClH83.383.327.8>99
14Cu-NPClH61.261.220.4>99
15Zn-NPBrH95.695.631.9>99
16Zn-NPIH96.396.332.1>99
17 eZn-NPClH70.8708141698
a Reaction conditions: PO (10 mL, 0.143 mol), Catalyst loading = 1 mol%, t = 3 h, T = 120 °C, P = 1 MPa. b Results determined by 1H NMR. c TON = n (product)/n (catalyst). d TOF = TON/reaction time. e Reaction conditions: PO (10 mL, 0.143 mol), Catalyst loading = 0.1 mol%, t = 0.5 h, T = 150 °C, P = 1 MPa.
Table 2. Cycloaddition between CO2 and various epoxides catalyzed by Zn-NPClH.
Table 2. Cycloaddition between CO2 and various epoxides catalyzed by Zn-NPClH.
EntryEpoxideProductCondition A aCondition B b
t (h)Yield (%)t (h)Yield (%)
1Materials 16 01646 i001Materials 16 01646 i002396--
2Materials 16 01646 i003Materials 16 01646 i004397--
3Materials 16 01646 i005Materials 16 01646 i0061032--
4Materials 16 01646 i007Materials 16 01646 i00810912499
5Materials 16 01646 i009Materials 16 01646 i01010222421
6Materials 16 01646 i011Materials 16 01646 i01210962488
7Materials 16 01646 i013Materials 16 01646 i0143942472
8Materials 16 01646 i015Materials 16 01646 i0163902490
9Materials 16 01646 i017Materials 16 01646 i0183882488
10Materials 16 01646 i019Materials 16 01646 i0203612468 c
11Materials 16 01646 i021Materials 16 01646 i0223522462
a Reaction conditions: 1 mol% Zn-NPClH, 0.143 mol epoxide; CO2 pressure 1 MPa; temperature 120 °C, the selectivity was determined by 1H NMR to be >99%. b Reaction conditions: 1 mol% Zn-NPClH, 0.143 mol epoxide; CO2 pressure 0.1 MPa; temperature 120 °C; the selectivity was determined by 1H NMR to be >99%. c Reaction temperature 110 °C.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wen, Q.; Yuan, X.; Zhou, Q.; Yang, H.-J.; Jiang, Q.; Hu, J.; Guo, C.-Y. Solvent-Free Coupling Reaction of Carbon Dioxide and Epoxides Catalyzed by Quaternary Ammonium Functionalized Schiff Base Metal Complexes under Mild Conditions. Materials 2023, 16, 1646. https://doi.org/10.3390/ma16041646

AMA Style

Wen Q, Yuan X, Zhou Q, Yang H-J, Jiang Q, Hu J, Guo C-Y. Solvent-Free Coupling Reaction of Carbon Dioxide and Epoxides Catalyzed by Quaternary Ammonium Functionalized Schiff Base Metal Complexes under Mild Conditions. Materials. 2023; 16(4):1646. https://doi.org/10.3390/ma16041646

Chicago/Turabian Style

Wen, Qin, Xuexin Yuan, Qiqi Zhou, Hai-Jian Yang, Qingqing Jiang, Juncheng Hu, and Cun-Yue Guo. 2023. "Solvent-Free Coupling Reaction of Carbon Dioxide and Epoxides Catalyzed by Quaternary Ammonium Functionalized Schiff Base Metal Complexes under Mild Conditions" Materials 16, no. 4: 1646. https://doi.org/10.3390/ma16041646

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop