Next Article in Journal
Direct Amide Synthesis over Composite Magnetic Catalysts in a Continuous Flow Reactor
Next Article in Special Issue
Recent Catalytic Advances in Hydrotreatment Processes of Pyrolysis Bio-Oil
Previous Article in Journal
C-,N- and S-Doped TiO2 Photocatalysts: A Review
Previous Article in Special Issue
CO Oxidation Efficiency and Hysteresis Behavior over Mesoporous Pd/SiO2 Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Syntheses, Characterization, and Application of Tridentate Phenoxyimino-Phenoxy Aluminum Complexes for the Coupling of Terminal Epoxide with CO2: From Binary System to Single Component Catalyst

1
School of Science, Shenyang University of Chemical Technology, Shenyang 110142, China
2
PolyAnalytik Inc., 700 Collip Circle, Suite 202, London, ON N6G 4X8, Canada
3
DUT Chemistry Analysis & Research Centre, Dalian University of Technology, Dalian 116024, China
*
Author to whom correspondence should be addressed.
Submission received: 30 December 2020 / Revised: 14 January 2021 / Accepted: 15 January 2021 / Published: 20 January 2021
(This article belongs to the Special Issue Sustainable and Environmental Catalysis)

Abstract

:
A series of binuclear aluminum complexes 13 supported by tridentate phenoxyimino-phenoxy ligands was synthesized and used as catalysts for the coupling reaction of terminal epoxide with carbon dioxide. The aluminum complex 1, which is catalytically inactive toward the coupling of epoxide with CO2 by itself, shows moderate activity in the presence of excess nucleophiles or organic bases at high temperature. In sharp contrast to complex 1, bifunctional complexes 2 and 3, which incorporate tertiary amine groups as the built-in nucleophile, are able to efficiently transform terminal epoxide with CO2 to corresponding cyclic carbonates as a sole product by themselves at 100 °C. The number of amine groups on the ligand skeleton and the reaction temperature exert a great influence on the catalytic activity. The bifunctional complexes 2 and 3 are also active at low carbon dioxide pressure such as 2 atm or atmospheric CO2 pressure. Kinetic studies of the coupling reactions of chloropropylene oxide/CO2 and styrene oxide/CO2 using bifunctional catalysts under atmospheric pressure of CO2 demonstrate that the coupling reaction has a first-order dependence on the concentration of the epoxide.

Graphical Abstract

1. Introduction

With the rapid development of the industry as well as the ever-growing human activity, a huge amount of CO2 has been released to the air, which has caused severe environmental problems [1,2,3,4]. The utilization of CO2 as feedstock to produce valuable chemicals is a promising way to solve this problem, but this conversion is limited due to the inertness of CO2 [5]. The catalytic coupling reaction with epoxides represents one of the promising processes employing CO2 to produce organic cyclic carbonate, which has found widespread applications in many respects [6,7,8,9,10]. Generally, the catalyst capable of achieving the transformation is comprised of nucleophile and Lewis acid. Thus, organic compounds and metal-based complexes have been employed as the Lewis acid for the coupling reaction, with the aid of nucleophiles to constitute the binary catalytic system. Usually, metal-based complexes display higher reactivity than organocatalysts for the coupling reaction. As a result of the diversity of the metal across the periodic table and the organic ligands of different structures, miscellaneous metal complexes based on aluminum [11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32], magnesium [33,34,35,36], chromium [37,38,39,40,41,42], cobalt [43,44,45,46,47,48], iron [49,50,51,52,53], and rare earth metals [54,55] have been innovated for the coupling reaction of CO2 with epoxides. Of many metal complexes, aluminum complexes have been drawing attention due to their low price, easy availability, and the superior selectivity for the cyclic carbonate over polymeric products [56]. It is well established that the electronic effects and steric hindrance around the aluminum center that can be finely tuned are crucial for the catalytic activity and the product selectivity. For example, optically active carbonate can be produced from the coupling of racemic epoxide with CO2 when the chiral complexes are employed [44,45,57].
In addition to binary catalytic system, there is a tendency to employ a bifunctional catalyst in which nucleophile and Lewis metal complexes are constructed into one molecule as the one-component catalyst for the coupling of CO2 with epoxides in the past decade [58,59,60,61,62,63,64,65,66,67,68]. Quaternary ammonium halide as a nucleophile is usually integrated into the metal complex to construct an ionic bifunctional catalyst. The pre-association of the nucleophilic halide to the metal complex via electrostatic attraction would effectively reduce the translational entropy penalty at the ring-opening step and the cyclization step as compared with the binary catalytic system [69]. Thus, the bifunctional catalyst loading for the coupling reaction is significantly reduced, and usually, a higher catalytic reactivity is achieved than that with the corresponding binary catalytic system. However, the incorporation of the quaternary onium salts into the metal complex impairs the solubility of the bifunctional catalyst. Another drawback is the thermostability of the ammonium halide-functionalized catalyst at elevated temperature [59], which undergoes ammonium salt decomposition pathways, including Zaitsev and Hoffman type eliminations [70,71,72] and retro-Menschutkin reactions [73,74,75,76]. Hence, it is important and necessary to develop bifunctional catalyst of high efficiency and stability at elevated temperature.
In this work, we present the syntheses of a series of aluminum complexes supported by novel tridentate phenoxyimino-phenoxy ligands and their catalytic behaviors for the coupling of terminal epoxides with CO2. Compared with tetradentate salen-type ligand stabilized metal complexes, which are extensively employed for the coupling of epoxide with CO2, aluminum complexes supported by tridentate ligands for this transformation are rare. As a result of the low coordination number of tridentate ligands, the resultant complexes are expected to have a more acidic metal center, which is supposed to benefit the activation of the epoxide ring and to facilitate the ring-opening. Moreover, an amine group was incorporated onto the ligand framework as the built-in nucleophile to construct neutral bifunctional complexes 2 and 3, in order to avoid the decomposition of ammonium halide moiety under elevated temperature. Systematic investigation of the catalytic behaviors of the aluminum complexes 2 and 3 reveals that the neutral and ionic bifunctional complexes are highly active toward the coupling of epoxide to CO2, whereas the unfunctionalized aluminum complex 1 is catalytically inactive by itself.

2. Results and Discussion

2.1. Syntheses and Characterization of Complexes 13

The tridentate ligand precursors H2L1–H2L3 (see Scheme 1) were readily synthesized by reacting the substituted salicylaldehyde with substituted aminophenol in 1:1 molar ratio (see the experimental section in the supporting material). The three ligand precursors were identified by proton nuclear magnetic resonance (1H NMR) and 13C-nuclear magnetic resonance (13C NMR) (Figures S1–S3 in the supporting material).
Subsequent syntheses of complexes 13 were conducted by the alkyl-elimination reaction of Et2AlCl with the corresponding ligand precursors at low temperature, as illustrated in Scheme 1. In a typical procedure, a tetrahydrofuran (THF) solution of H2L1 was added dropwise to a stirred hexane solution of AlEt2Cl at −30 °C to afford a yellow solution. The reaction was slowly warmed up to room temperature and stirred for another 4 h. After the removal of all the volatiles under reduced pressure, yellow solids were separated, which was followed by washing with cold hexane three times and drying to constant weight under vacuum. The structure of the isolated solids was characterized by the NMR technique with CDCl3 as the solvent. The 1H NMR spectrum shows no resonances of hydroxy groups or ethyl groups bound to aluminum. In the low field, the resonance of 8.06 ppm assignable to the imino moiety was observed, which slightly shifted to the low field when compared with that in free ligand precursor H2L1 (see Figure S4 in the supporting material). Two distinct resonances of 1.32 ppm and 0.84 ppm for the tBu groups were found (1.49 and 1.35 ppm in the H2L1), indicating chelation of the tridentate ligand to aluminum. Moreover, the upfield shift of the imino group in the H2L1 from 8.72 to 8.06 ppm in the isolated complex suggested the coordination of nitrogen of the imino group to the aluminum center. Remarkably, no solvent molecules were detected in the 1H NMR spectrum, suggesting that the isolated complex is THF-free.
In addition to the characterization by NMR, mass spectrometry (MS) was employed to characterize the isolated complex. A very weak peak at m/z = 350.1566 was detected, which corresponds to the cationic species 1a having the formula of C21H25O2NAl, as shown in Scheme 2. Interestingly, a strong signal was observed at m/z = 414.2219, which was assigned to the cationic species 1b with the formula of C23H33O4NAl. We supposed that the species 1b was generated by the coordination of two methanol molecules to the cationic species 1a. Thus, it is reasonable to interpret the weak signal of 1a and the strong peak of 1b on the mass spectrum. In the light of the strong signal of 1b on the MS, no solvent detection on the 1H NMR spectrum, and the coordination number of 5 usually adopted by aluminum complex, we expected that complex 1 should be a binuclear structure, with each aluminum center being surrounded by a phenoxyimino-phenoxy ligand and two chlorides, as presented in Scheme 2.
Similarly, treatment of the ligand precursors H2L2 and H2L3 with AlEt2Cl under low temperature resulted in the formation of bifunctional complexes 2 and 3. The characterization of complex 2 by 1H NMR showed the successful alkyl elimination between AlEt2Cl and H2L2. Moreover, there are no THF molecules detected on the 1H NMR spectrum, suggesting that complex 2 is THF-free. The MS spectrum of complex 2 shows a m/z peak positioning at 397.1085. We tentatively ascribed this signal to the structure 2a, which probably is an ethanol coordinated species, as illustrated in Figure 1. Meanwhile, a peak at m/z = 783.3226 was detected, which is ascribed to the cationic species 2b. Although a tertiary amine group is introduced into the molecule, the inner coordination of the amine group to the Al center is impossible because of the orientation of the tertiary amine group. Taking into account the coordination saturation of the aluminum, it is reasonable to assume the complex 2 adopts binuclear structure with dangling tertiary amine groups. Similar results were also observed for complex 3, as indicated by the 1H NMR and MS, which should also adopt a binuclear structure with dangling amine groups via chloride bridges.

2.2. Catalytic Performance of Complexes 13 in the Coupling Reactions of Terminal Epoxide with CO2

To begin with, we examined the catalytic performance of complex 1 for the coupling of propylene oxide (PO) with CO2 (see Scheme 3). Unfortunately, complex 1 was catalytically inactive by itself, which was in agreement with the previous literature reports [21,22]. When it was paired with a nucleophile or organic base, complex 1 showed moderate activity for the PO/CO2 coupling (see Tables S1–S4 in the supporting information). For instance, complex 1 in combination with tetrabutylammonium bromide (TBAB, 40 eq) catalyzed the PO/CO2 coupling to produce propylene carbonate (PC) as the sole product at 80 °C under 20 atm CO2 in 10 h with 91% PO conversion (see entry 8 in Table S1). There was no detection of any polymeric products as evidenced by the 1H NMR spectrum of the reaction mixtures. It was also interesting to find that the complex 1/TBAB (40 eq) system displayed a zeroth-order dependence on the PO throughout the coupling (see Table S2 and Figure S7, in the supporting material). This result is analogous to Han’s report in which polymeric ionic liquid was employed as the catalyst for the PO/CO2 coupling reaction [77]. When other organobases such as 4-dimethylaminopyridine (DMAP) or 1-methylimidazole (1-MeIm) were employed as the cocatalysts, high cocatalyst loading was required for effective transformation (see Tables S3 and S4 in the supporting material). For instance, the binary system composed of complex 1/DMAP (1/20) displayed lower catalytic activity than complex 1/TBAB (1/20) at 60 °C.
It has been reported by North and Pasquale that quaternary ammonium halide would decompose at high temperature to generate tertiary amine and halohydrocarbon [78]. The in situ generated tertiary amine, together with the remaining ammonium halide salts, assists the metal complex to convert propylene oxide and CO2 to cyclic carbonate. Recently, Kim and his coworkers reported the coupling of epoxide with CO2 catalyzed by a series of tertiary amines with moderate catalytic activity [79]. These results enlightened us to design amine-functionalized complex 2 as the single-component catalyst for the coupling of epoxide with CO2. The catalytic results were compiled in Table 1. It was satisfying that the bifunctional complex 2 by itself succeeded in transforming the CO2 and PO to PC. Analysis of the reaction mixture showed that PC was exclusively produced without the formation of any polymeric products as evidenced by 1H NMR spectrum. Although complex 2 displayed very low activity at 80 °C (entries 1 and 2, in Table 1), however, the catalytic activity was considerably enhanced when the reaction was conducted at 100 °C. For instance, 68% of PO was converted to PC under 10 atm CO2 in 10 h (entry 3, Table 1). In addition to the reaction temperature, increasing the CO2 pressure accelerates the coupling reaction. Then, 94% of PO was transformed to PC at 20 atm CO2 pressure within 10 h (entry 4, Table 1), suggesting that high CO2 pressure, namely, the high CO2 concentration facilitated the coupling reaction.
Subsequent examination of the catalytic performance of complex 2 in other terminal epoxides/CO2 coupling reactions confirmed again the significance of the reaction temperature and CO2 pressure. The 1,2-butylene oxide (BO) conversion of 7% at 80 °C under 10 atm CO2 was improved to 49% when the reaction was performed at 100 °C under 20 atm of CO2 pressure (entries 5 and 6, Table 1). Similar trends were also observed for the coupling of CO2 with epichlorohydrin (or chloropropylene oxide, abbreviated as CPO) or styrene oxide (SO). The CPO conversion of 12% at 80 °C under 10 atm of CO2 in 10 h was enhanced to 36%, 52%, and 90% within 5, 10, and 24 h, respectively, when the reaction temperature was simply increased to 100 °C (entries 7, 9–11, Table 1). As the CO2 pressure was increased to 20 atm, the conversion of CPO rose to 40% in 5 h and 62% in 10 h under 20 atm CO2 (entries 12 and 13, Table 1). As for the SO/CO2 coupling, the reaction was quite sluggish (entries 14–19, Table 1). High SO conversion can only be achieved after a long reaction time (entry 17, Table 1).
Complex 2 was capable of transforming CPO/CO2 and SO/CO2 to cyclic carbonate under atmospheric CO2 pressure. The results were collected in Table 2, and the kinetic curves (conversion vs. reaction time) were depicted in Figure 2. As expected, the rate of the coupling of CO2 with either CPO or SO under 1 atm CO2 was extremely slow, which was partly due to the low CO2 pressure. As can be seen from Figure 2, there is a surge in the CPO conversion after 36 h (the blue curve). This might be interpreted by the poor solubility of complex 2 in CPO. Therefore, the metal complex 2 was not well dispersed in the epoxide, leading to a heterogeneous system. Thus, the effective concentration of complex 2 was lower than the theoretical value at the beginning. With the proceeding of the coupling reaction, highly polar cyclic carbonate was generated, which in turn acted as solvent to dissolve complex 2. Thus, the effective concentration of complex 2 was continuously increased in the course of the reaction until the full dissolution of complex 2. This would accelerate the coupling, as is shown in Figure 2 (the blue curve). A similar situation took place in the case of SO/CO2 coupling as well, where the reaction rate gradually increased with the reaction time, as indicated by the kinetic curve (the red curves in Figure 2).
The success of complex 2 as a single-component catalyst motivated us to examine the catalytic performance of complex 3, which has more tertiary amine groups on the ligand skeleton for the coupling of epoxides toward CO2. The catalytic results were collected in Table 3. As anticipated, complex 3 displayed higher activity than complex 2 did under the same conditions, which was tentatively attributed to the incorporation of one more amine group. Similar to complex 2, complex 3 exhibited low activity at 80 °C. For example, 14% of PO was transformed to PC under 20 atm CO2 in 10 h (entry 1, Table 3). However, the catalytic activity was dramatically improved by increasing the temperature to 100 °C, with the 64% conversion of PO in 5 h, and 99% in 10 h (entries 2 and 3, Table 3). The significance of reaction temperature was also observed in the BO/CO2 coupling. The BO conversion of 5% at 80 °C was dramatically increased to 76% when the reaction temperature was increased to 100 °C at 10 atm CO2 pressure (entries 4 and 5, Table 3). Similarly, enhancing the CO2 pressure to 20 atm gave rise to 95% of BO conversion in 10 h (entry 6, Table 3), which was much higher than that of the coupling reaction catalyzed by complex 2 (entry 6, Table 1).
Complex 3 displayed high catalytic activity for the CPO/CO2 and SO/CO2 coupling. Taking the CPO/CO2 coupling as an example, complex 3 converted 92% of CPO in 5 h at 100 °C under 20 atm of CO2 to produce the corresponding carbonate (entry 7, Table 3), and a higher conversion rate of 98% was achieved in 10 h (entry 8, Table 3). When SO was used as the substrate, 50% of the SO was converted to the corresponding cyclic carbonate in 10 h at 100 °C under 20 atm of CO2 (entry 9, Table 3). Nearly complete conversion of SO was achieved by extension of the reaction time to 24 h (entry 10, Table 3).
Complex 3 also displayed high activity toward CPO/CO2 coupling under low CO2 pressure. For example, 85% of CPO was converted to cyclic carbonate at 100 °C within 5 h under constant 2 atm CO2 pressure (entry 1, Table 4). Prolonging the reaction time to 10 h gave rise to the nearly complete conversion of CPO (entry 2, Table 4). Complex 3 was still active under atmospheric CO2 pressure, but the activity dropped sharply. For example, the conversion of CPO was only 28% in 5 h and 55% in 10 h (entries 1 and 2, Table 4), owing to the low concentration of CO2 dissolved in CPO. By further extending the reaction time, high conversion of CPO was achieved (entries 3–6, Table 4). However, for SO/CO2 coupling, complex 3 displayed very low activity. The conversion of SO was only 26% in 10 h when the reaction was performed at 100 °C, maintaining the CO2 pressure at constant 2 atm (entry 7, Table 4). Lower SO conversion was expected, as the reaction was carried out at atmospheric CO2 pressure. Analogous to CPO/CO2 coupling, a high conversion of SO can be achieved by extending the reaction time (entries 8–14, Table 4).
Kinetic studies of the coupling reactions of CPO/CO2 and SO/CO2 at atmospheric CO2 pressure by complex 3 are plotted in Figure 3. It is clear that the CPO/CO2 coupling is faster than the SO/CO2 coupling under the same conditions (the blue curves in Figure 3). It is worth noting that the coupling of CPO/CO2 has a first-order dependence on CPO concentration throughout the coupling reaction, as illustrated in Figure 3 (the red dash line, -ln(1-Conversion) vs. time). This may indicate a better solubility of complex 3 in CPO than that of complex 2, which was tentatively ascribed to the incorporation of more amine groups on the ligand framework. Differing from CPO/CO2 coupling, the kinetic curve (-ln(1-Conversion) vs. time) of SO/CO2 coupling shows an increase in the slope during the early stage of the reaction (the red dot curve in Figure 3), reflecting the heterogeneous nature of the system owing to the poor solubility of complex 3 in SO. As the reaction proceeded, more SC was produced, which in turn helped to dissolve complex 3, increasing the catalyst concentration and speeding up the coupling reaction. After some time, complex 3 was totally dissolved, forming a homogeneous system and maintaining constant concentration. Thus, a linear relationship was observed at the late stage of the coupling reaction, demonstrating a first-order dependence on SO concentration. The linear relation at the late stage also suggested the high stability of complex 3 under elevated temperature such as 100 °C.

3. Materials and Methods

2-aminophenol (99%), 4-tertbutylphenol, aqueous dimethylamine solution (40 wt%), 3,5-ditertbutylsalicyaldehyde (98%), and diethyl aluminum chloride (1 mol/L, in hexane) were purchased from Energy Chemical (Anqing, China) and used as received. Propylene oxide, 1,2-butylene oxide, epichlorohydrin, and styrene oxide obtained from Energy Chemical (Anqing, China) were distilled over CaH2 before use. THF and hexane were pre-dried over 4 Å molecular sieves before distillation over sodium with benzophenone as the indicator under an argon atmosphere and stored over freshly cut sodium in a glovebox. CO2 (99.999% purity) was purchased from Shenyang Hongsheng Gas Limited Corporation, Suzhou, China. Ethanol and formic acid were used as received without further handling.
General procedures: All reactions sensitive to air and moisture were carried out in a glovebox filled with dry argon. Proligands H2L1-H2L3 were synthesized according to the literature methods and structurally identified by 1H NMR and 13C NMR [24]. The aluminum complexes 13 were prepared by equimolar reaction of the proligands with AlEt2Cl in a glovebox filled with argon. The structures of complexes 13 were characterized by 1H NMR, 13C NMR, and MS. The coupling of epoxide with CO2 was carried in a stainless-steel autoclave equipped with a magnetic stirring bar. The reaction mixture was analyzed by the 1H NMR spectra with CDCl3 as a solvent. The 1H and 13C NMR spectra were recorded using Bruker AVANCE III 500 MHz spectrometer (Billerica, MA, USA). Mass spectra of complexes 13 were obtained in the electrospray positive mode (ESI+) on Thermo LTQ Orbitrap XL spectrometer (Waltham, MA, USA), samples were diluted in methanol or ethanol, at DUT Chemistry Analysis & Research Centre, Dalian University of Technology (Dalian, China).

3.1. Coupling Reaction of Epoxide with CO2 by Aluminum Complex 1

The typical procedure for the coupling of PO with CO2 by complex 1 under elevated pressure is as follows. Complex 1 and cocatalyst were dissolved in PO in a Schlenk tube. Then, the solution was transferred via syringe into the pre-dried autoclave under CO2 atmosphere. Then, the autoclave was pressurized with CO2 and heated. After the designated time, the autoclave was cooled in an ice bath. The excess of the CO2 was vented out. The conversion of PO was determined by GC analysis and 1H NMR of the reaction mixture, and the yield was calculated based on isolated propylene carbonate. The results showed that the GC result was consistent with that by weight analysis of the PC.

3.2. Coupling Reaction of Epoxide with CO2 by Bifunctional Aluminum Complexes

The typical procedure for the coupling of PO with CO2 by complex 2 or 3 under elevated pressure is as follows. Complex 2 (38.7 mg, 50 μmol) was first placed in the pre-dried autoclave. After the internal atmosphere of the autoclave was displaced by CO2 three times, PO was injected, which was followed by charging CO2 to 20 atm. Then, the autoclave was heated for the designated time. The analysis was the same as the above method.
The typical procedure for the coupling of CO2 with CPO by complex 2 or 3 at atmospheric CO2 pressure is as follows. Complex 2 (38.7 mg, 50 μmol) was placed in a 10 mL round-bottom flask in a glovebox. The flask was taken out, which was equipped with a balloon. The flask was vacuumed and recharged with CO2. After the injection of a prescribed amount of CPO, the flask was heated and maintained at 100 °C. The samples were taken out periodically and analyzed by Agilent GC to determine the conversion of CPO.

4. Conclusions

We have demonstrated the synthesis of a series of aluminum complexes 13. Structural characterization by NMR and MS suggested the formation of binuclear structures via the chloride. Although complex 1 can be activated by either a nucleophile or an organic base to enable the coupling of terminal epoxide with CO2, the excess use of cocatalyst unambiguously hampers its practical application. In contrast, the bifunctional complexes 2 and 3 containing tertiary amine as the built-in nucleophile are highly active catalysts by their own for the coupling of CO2 with terminal epoxides. For bifunctional complexes 2 and 3, high temperature facilitates the rapid transformation. It is found that the incorporation of more amine groups in the metal complex greatly enhances the catalytic activity. The bifunctional complexes are even active under low CO2 pressure at the expense of catalytic activity to some extent. This work may shed light on the design of metal-based catalyst of high activity for the coupling of terminal epoxide with CO2.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4344/11/2/145/s1, Figure S1: 1H NMR spectrum of proligand H2L1 in CDCl3, Figure S1: 1H NMR spectrum of proligand H2L2 in DMSO-d6, Figure S3: 1H NMR spectrum of proligand H2L3 in DMSO-d6, Figure S4: 1H NMR spectrum of the complex 1 in CDCl3, Figure S5: 1H NMR spectrum of the complex 2 in DMSO-d6, Figure S6: 1H NMR spectrum of the complex 3 in DMSO-d6, Figure S7: Kinetic study of coupling of PO/CO2 mediated by 1/TBAB (1:40 molar ratio), PO conversion vs. reaction time, Table S1: Cycloaddition of CO2 to PO mediated by complex 1 and TBAB, Table S2: Effect of reaction time on the PO conversion, Table S3: Coupling of CO2/PO catalyzed by 1/DMAP, Table S4: Coupling of CO2/PO catalyzed by 1/1-MeIm.

Author Contributions

Conceptualization, Z.Z. and S.H.; validation, S.H., Q.D., T.W. and Z.Z.; investigation, Z.Z., Q.D. and T.W.; writing—original draft preparation, Z.Z.; writing—review and editing, Q.D., P.X. and S.H.; funding acquisition, Z.Z. and S.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by “National Natural Science Foundation of China, grant number 21674066”, and “Ministry of Education of Liaoning province, China, grant number LJ2020009, LQ2020028”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article or supplementary material.

Acknowledgments

We thank Meiheng Lv for her constructive discussions about the identification of bifunctional complexes 2 and 3.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dibenedetto, A.; Angelini, A.; Stufano, P. Use of carbon dioxide as feedstock for chemicals and fuels: Homogeneous and heterogeneous catalysis. J. Chem. Technol. Biotechnol. 2014, 89, 334–353. [Google Scholar] [CrossRef]
  2. Yu, B.; Diao, Z.-F.; Guo, C.-X.; He, L.-N. Carboxylation of olefins/alkynes with CO2 to industrially relevant acrylic acid derivatives. J. CO2 Util. 2013, 1, 60–68. [Google Scholar] [CrossRef]
  3. Drees, M.; Cokoja, M.; Kühn, F.E. Recycling CO2? Computational Considerations of the Activation of CO2 with Homogeneous Transition Metal Catalysts. ChemCatChem 2012, 4, 1703–1712. [Google Scholar] [CrossRef]
  4. Cokoja, M.; Bruckmeier, C.; Rieger, B.; Herrmann, W.A.; Kuehn, F.E. Transformation of Carbon Dioxide with Homogeneous Transition-Metal Catalysts: A Molecular Solution to a Global Challenge? Angew. Chem. Int. Ed. 2011, 50, 8510–8537. [Google Scholar] [CrossRef] [PubMed]
  5. Omae, I. Aspects of carbon dioxide utilization. Catal. Today 2006, 115, 33–52. [Google Scholar] [CrossRef]
  6. Cokoja, M.; Wilhelm, M.E.; Anthofer, M.H.; Herrmann, W.A.; Kuehn, F.E. Synthesis of Cyclic Carbonates from Epoxides and Carbon Dioxide by Using Organocatalysts. ChemSusChem 2015, 8, 2436–2454. [Google Scholar] [CrossRef] [PubMed]
  7. Maeda, C.; Miyazaki, Y.; Ema, T. Recent progress in catalytic conversions of carbon dioxide. Catal. Sci. Technol. 2014, 4, 1482–1497. [Google Scholar] [CrossRef] [Green Version]
  8. Xu, K. Nonaqueous liquid electrolytes for lithium-based rechargeable batteries. Chem. Rev. 2004, 104, 4303–4417. [Google Scholar] [CrossRef]
  9. Schaeffner, B.; Schaeffner, F.; Verevkin, S.P.; Boerner, A. Organic Carbonates as Solvents in Synthesis and Catalysis. Chem. Rev. 2010, 110, 4554–4581. [Google Scholar] [CrossRef]
  10. Shaikh, A.A.G.; Sivaram, S. Organic carbonates. Chem. Rev. 1996, 96, 951–976. [Google Scholar] [CrossRef]
  11. Lu, X.B.; Feng, X.J.; He, R. Catalytic formation of ethylene carbonate from supercritical carbon dioxide/ethylene oxide mixture with tetradentate Schiff-base complexes as catalyst. Appl. Catal. Gen. 2002, 234, 25–33. [Google Scholar] [CrossRef]
  12. Fuchs, M.A.; Altesleben, C.; Zevaco, T.A.; Dinjus, E. An Efficient Homogeneous Chloro-Aluminum- N2O2 Catalyst for the Coupling of Epoxides with Carbon Dioxide. Eur. J. Inorg. Chem. 2013, 2013, 4541–4545. [Google Scholar] [CrossRef]
  13. Clegg, W.; Harrington, R.W.; North, M.; Pasquale, R. Cyclic Carbonate Synthesis Catalysed by Bimetallic Aluminium-Salen Complexes. Chem. Eur. J. 2010, 16, 6828–6843. [Google Scholar] [CrossRef] [PubMed]
  14. Kasuga, K.; Nagao, S.; Fukumoto, T.; Handa, M. Cycloaddition of carbon dioxide to propylene oxide catalysed by tetra-t-butylphthalocyaninatoaluminium(III) chloride. Polyhedron 1996, 15, 69–72. [Google Scholar] [CrossRef]
  15. Woo, W.H.; Hyun, K.; Kim, Y.; Ryu, J.Y.; Lee, J.; Kim, M.; Park, M.H.; Kim, Y. Highly Active Salen-Based Aluminum Catalyst for the Coupling of Carbon Dioxide with Epoxides at Ambient Temperature. Eur. J. Inorg. Chem. 2017, 5372–5378. [Google Scholar] [CrossRef]
  16. Wu, X.; North, M. A Bimetallic Aluminium(Salphen) Complex for the Synthesis of Cyclic Carbonates from Epoxides and Carbon Dioxide. ChemSusChem 2017, 10, 74–78. [Google Scholar] [CrossRef] [PubMed]
  17. Li, C.-Y.; Liu, D.-C.; Ko, B.-T. Synthesis, characterization and reactivity of single-site aluminium amides bearing benzotriazole phenoxide ligands: Catalysis for ring-opening polymerization of lactide and carbon dioxide/propylene oxide coupling. Dalton Trans. 2013, 42, 11488–11496. [Google Scholar] [CrossRef]
  18. Melendez, J.; North, M.; Pasquale, R. Synthesis of cyclic carbonates from atmospheric pressure carbon dioxide using exceptionally active aluminium(salen) complexes as catalysts. Eur. J. Inorg. Chem. 2007, 3323–3326. [Google Scholar] [CrossRef]
  19. Qin, Y.; Guo, H.; Sheng, X.; Wang, X.; Wang, F. An aluminum porphyrin complex with high activity and selectivity for cyclic carbonate synthesis. Green Chem. 2015, 17, 2853–2858. [Google Scholar] [CrossRef]
  20. Aida, T.; Inoue, S. Activation of carbon dioxide with aluminum porphyrin and reaction with epoxide. Studies on (tetraphenylporphinato)aluminum alkoxide having a long oxyalkylene chain as the alkoxide group. J. Am. Chem. Soc. 1983, 105, 1304–1309. [Google Scholar] [CrossRef]
  21. Lu, X.B.; Zhang, Y.J.; Jin, K.; Luo, L.M.; Wang, H. Highly active electrophile-nucleophile catalyst system for the cycloaddition of CO2 to epoxides at ambient temperature. J. Catal. 2004, 227, 537–541. [Google Scholar] [CrossRef]
  22. Lu, X.B.; Zhang, Y.J.; Liang, B.; Li, X.; Wang, H. Chemical fixation of carbon dioxide to cyclic carbonates under extremely mild conditions with highly active bifunctional catalysts. J. Mol. Catal. Chem. 2004, 210, 31–34. [Google Scholar] [CrossRef]
  23. Rintjema, J.; Kleij, A.W. Aluminum-Mediated Formation of Cyclic Carbonates: Benchmarking Catalytic Performance Metrics. ChemSusChem 2017, 10, 1274–1282. [Google Scholar] [CrossRef] [PubMed]
  24. Gao, P.; Zhao, Z.; Chen, L.; Yuan, D.; Yao, Y. Dinuclear Aluminum Poly(phenolate) Complexes as Efficient Catalysts for Cyclic Carbonate Synthesis. Organometallics 2016, 35, 1707–1712. [Google Scholar] [CrossRef]
  25. Cozzolino, M.; Press, K.; Mazzeo, M.; Lamberti, M. Carbon Dioxide/Epoxide Reactions Catalyzed by Bimetallic Salalen Aluminum Complexes. ChemCatChem 2016, 8, 455–460. [Google Scholar] [CrossRef]
  26. Castro-Osma, J.A.; Lara-Sanchez, A.; North, M.; Otero, A.; Villuendas, P. Synthesis of cyclic carbonates using monometallic, and helical bimetallic, aluminium complexes. Catal. Sci. Technol. 2012, 2, 1021–1026. [Google Scholar] [CrossRef]
  27. Rios Yepes, Y.; Quintero, C.; Osorio Melendez, D.; Daniliuc, C.G.; Martinez, J.; Rojas, R.S. Cyclic Carbonates from CO2 and Epoxides Catalyzed by Tetra- and Pentacoordinate Amidinate Aluminum Complexes. Organometallics 2019, 38, 469–478. [Google Scholar] [CrossRef]
  28. Whiteoak, C.J.; Kielland, N.; Laserna, V.; Castro-Gomez, F.; Martin, E.; Escudero-Adan, E.C.; Bo, C.; Kleij, A.W. Highly Active Aluminium Catalysts for the Formation of Organic Carbonates from CO2 and Oxiranes. Chem. Eur. J. 2014, 20, 2264–2275. [Google Scholar] [CrossRef]
  29. Castro-Osma, J.A.; North, M.; Wu, X. Development of a Halide-Free Aluminium-Based Catalyst for the Synthesis of Cyclic Carbonates from Epoxides and Carbon Dioxide. Chem. Eur. J. 2014, 20, 15005–15008. [Google Scholar] [CrossRef]
  30. Supasitmongkol, S.; Styring, P. A single centre aluminium(III) catalyst and TBAB as an ionic organo-catalyst for the homogeneous catalytic synthesis of styrene carbonate. Catal. Sci. Technol. 2014, 4, 1622–1630. [Google Scholar] [CrossRef] [Green Version]
  31. Kim, S.H.; Han, S.Y.; Kim, J.H.; Kang, Y.Y.; Lee, J.; Kim, Y. Monomeric or Dimeric Aluminum Complexes as Catalysts for Cycloaddition between CO2 and Epoxides. Eur. J. Inorg. Chem. 2015, 2015, 2323–2329. [Google Scholar] [CrossRef]
  32. Whiteoak, C.J.; Kielland, N.; Laserna, V.; Escudero-Adan, E.C.; Martin, E.; Kleij, A.W. A Powerful Aluminum Catalyst for the Synthesis of Highly Functional Organic Carbonates. J. Am. Chem. Soc. 2013, 135, 1228–1231. [Google Scholar] [CrossRef] [PubMed]
  33. Ema, T.; Miyazaki, Y.; Koyama, S.; Yano, Y.; Sakai, T. A bifunctional catalyst for carbon dioxide fixation: Cooperative double activation of epoxides for the synthesis of cyclic carbonates. Chem. Commun. 2012, 48, 4489–4491. [Google Scholar] [CrossRef] [PubMed]
  34. Bai, D.; Duan, S.; Hai, L.; Jing, H. Carbon Dioxide Fixation by Cycloaddition with Epoxides, Catalyzed by Biomimetic Metalloporphyrins. ChemCatChem 2012, 4, 1752–1758. [Google Scholar] [CrossRef]
  35. Raghavendra, B.; Shashank, P.V.S.; Pandey, M.K.; Reddy, N.D. CO2/Epoxide Coupling and the ROP of epsilon-Caprolactone: Mg and Al Complexes of gamma-Phosphino-ketiminates as Dual-Purpose Catalysts. Organometallics 2018, 37, 1656–1664. [Google Scholar] [CrossRef]
  36. 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] [CrossRef]
  37. Darensbourg, D.J.; Fang, C.C.; Rodgers, J.L. Catalytic coupling of carbon dioxide and 2,3-epoxy-1,2,3,4-tetrahydronaphthalene in the presence of a (Salen)(CrCl)-Cl-III derivative. Organometallics 2004, 23, 924–927. [Google Scholar] [CrossRef]
  38. Castro-Osma, J.A.; North, M.; Wu, X. Synthesis of Cyclic Carbonates Catalysed by Chromium and Aluminium Salphen Complexes. Chem. Eur. J. 2016, 22, 2100–2107. [Google Scholar] [CrossRef]
  39. Ramidi, P.; Sullivan, S.Z.; Gartia, Y.; Munshi, P.; Griffin, W.O.; Darsey, J.A.; Biswas, A.; Shaikh, A.U.; Ghosh, A. Catalytic Cyclic Carbonate Synthesis Using Epoxide and Carbon Dioxide: Combined Catalytic Effect of Both Cation and Anion of an Ionic Cr-v(O) Amido Macrocyclic Complex. Ind. Eng. Chem. Res. 2011, 50, 7800–7807. [Google Scholar] [CrossRef]
  40. Cuesta-Aluja, L.; Djoufak, M.; Aghmiz, A.; Rivas, R.; Christ, L.; Masdeu-Bulto, A.M. Novel chromium (III) complexes with N-4-donor ligands as catalysts for the coupling of CO2 and epoxides in supercritical CO2. J. Mol. Catal. Chem. 2014, 381, 161–170. [Google Scholar] [CrossRef]
  41. Adolph, M.; Zevaco, T.A.; Altesleben, C.; Walter, O.; Dinjus, E. New cobalt, iron and chromium catalysts based on easy-to-handle N-4-chelating ligands for the coupling reaction of epoxides with CO2. Dalton Trans. 2014, 43, 3285–3296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Iksi, S.; Aghmiz, A.; Rivas, R.; Dolores Gonzalez, M.; Cuesta-Aluja, L.; Castilla, J.; Orejon, A.; El Guemmout, F.; Masdeu-Bulto, A.M. Chromium complexes with tridentate NN′O Schiff base ligands as catalysts for the coupling of CO2 and epoxides. J. Mol. Catal. Chem. 2014, 383, 143–152. [Google Scholar] [CrossRef]
  43. Ambrose, K.; Robertson, K.N.; Kozak, C.M. Cobalt amino-bis(phenolate) complexes for coupling and copolymerization of epoxides with carbon dioxide. Dalton Trans. 2019, 48, 6248–6260. [Google Scholar] [CrossRef]
  44. Lu, X.B.; Liang, B.; Zhang, Y.J.; Tian, Y.Z.; Wang, Y.M.; Bai, C.X.; Wang, H.; Zhang, R. Asymmetric catalysis with CO2: Direct synthesis of optically active propylene carbonate from racemic epoxides. J. Am. Chem. Soc. 2004, 126, 3732–3733. [Google Scholar] [CrossRef] [PubMed]
  45. Chang, T.; Jin, L.; Jing, H. Bifunctional Chiral Catalyst for the Synthesis of Chiral Cyclic Carbonates from Carbon Dioxide and Epoxides. ChemCatChem 2009, 1, 379–383. [Google Scholar] [CrossRef]
  46. Reiter, M.; Altenbuchner, P.T.; Kissling, S.; Herdtweck, E.; Rieger, B. Amine-bis(phenolato)cobalt(II) Catalysts for the Formation of Organic Carbonates from Carbon Dioxide and Epoxides. Eur. J. Inorg. Chem. 2015, 2015, 1766–1774. [Google Scholar] [CrossRef]
  47. Yu, C.-Y.; Chuang, H.-J.; Ko, B.-T. Bimetallic bis(benzotriazole iminophenolate) cobalt, nickel and zinc complexes as versatile catalysts for coupling of carbon dioxide with epoxides and copolymerization of phthalic anhydride with cyclohexene oxide. Catal. Sci. Technol. 2016, 6, 1779–1791. [Google Scholar] [CrossRef]
  48. Adolph, M.; Zevaco, T.A.; Altesleben, C.; Staudt, S.; Walter, O.; Dinjus, E. New ionic cobalt(III) complexes based on the N, N-bis(2-pyrazinecarboxamide)-1,2-benzene ligand: Application to the formation of organic carbonates from epoxides and CO2. New J. Chem. 2015, 39, 9858–9865. [Google Scholar] [CrossRef]
  49. Buonerba, A.; De Nisi, A.; Grassi, A.; Milione, S.; Capacchione, C.; Vagin, S.; Rieger, B. Novel iron(III) catalyst for the efficient and selective coupling of carbon dioxide and epoxides to form cyclic carbonates. Catal. Sci. Technol. 2015, 5, 118–123. [Google Scholar] [CrossRef]
  50. Buchard, A.; Kember, M.R.; Sandeman, K.G.; Williams, C.K. A bimetallic iron(III) catalyst for CO2/epoxide coupling. Chem. Commun. 2011, 47, 212–214. [Google Scholar] [CrossRef] [Green Version]
  51. Whiteoak, C.J.; Martin, E.; Martinez Belmonte, M.; Benet-Buchholz, J.; Kleij, A.W. An Efficient Iron Catalyst for the Synthesis of Five- and Six-Membered Organic Carbonates under Mild Conditions. Adv. Synth. Catal. 2012, 354, 469–476. [Google Scholar] [CrossRef]
  52. Della Monica, F.; Buonerba, A.; Capacchione, C. Homogeneous Iron Catalysts in the Reaction of Epoxides with Carbon Dioxide. Adv. Synth. Catal. 2019, 361, 265–282. [Google Scholar] [CrossRef]
  53. Della Monica, F.; Vummaleti, S.V.C.; Buonerba, A.; De Nisi, A.; Monari, M.; Milione, S.; Grassi, A.; Cavallo, L.; Capacchione, C. Coupling of Carbon Dioxide with Epoxides Efficiently Catalyzed by Thioether-Triphenolate Bimetallic Iron(III) Complexes: Catalyst Structure-Reactivity Relationship and Mechanistic DFT Study. Adv. Synth. Catal. 2016, 358, 3231–3243. [Google Scholar] [CrossRef]
  54. Zhao, Z.; Qin, J.; Zhang, C.; Wang, Y.; Yuan, D.; Yao, Y. Recyclable Single-Component Rare-Earth Metal Catalysts for Cycloaddition of CO2 and Epoxides at Atmospheric Pressure. Inorg. Chem. 2017, 56, 4568–4575. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, C.; Liu, X.; Dai, Z.; Sun, Y.; Tang, N.; Wu, J. Yttrium complex supported by a sterically encumbering N-anchored tris-arylphenoxide ligand: Heteroselective ROP of rac-lactide and CO2/epoxide coupling. Inorg. Chem. Commun. 2015, 56, 69–72. [Google Scholar] [CrossRef]
  56. Luinstra, G.A.; Haas, G.R.; Molnar, F.; Bernhart, V.; Eberhardt, R.; Rieger, B. On the formation of aliphatic polycarbonates from epoxides with chromium(III) and aluminum(III) metal-salen complexes. Chem. Eur. J. 2005, 11, 6298–6314. [Google Scholar] [CrossRef]
  57. North, M.; Quek, S.C.Z.; Pridmore, N.E.; Whitwood, A.C.; Wu, X. Aluminum(salen) Complexes as Catalysts for of Terminal Epoxides via CO2 Coupling. ACS Catal. 2015, 5, 3398–3402. [Google Scholar] [CrossRef]
  58. North, M.; Young, C. Bimetallic aluminium(acen) complexes as catalysts for the synthesis of cyclic carbonates from carbon dioxide and epoxides. Catal. Sci. Technol. 2011, 1, 93–99. [Google Scholar] [CrossRef]
  59. Rulev, Y.A.; Gugkaeva, Z.; Maleev, V.I.; North, M.; Belokon, Y.N. Robust bifunctional aluminium-salen catalysts for the preparation of cyclic carbonates from carbon dioxide and epoxides. Beilstein. J. Org. Chem. 2015, 11, 1614–1623. [Google Scholar] [CrossRef] [Green Version]
  60. Martinez, J.; Castro-Osma, J.A.; Alonso-Moreno, C.; Rodriguez-Dieguez, A.; North, M.; Otero, A.; Lara-Sanchez, A. One-Component Aluminum(heteroscorpionate) Catalysts for the Formation of Cyclic Carbonates from Epoxides and Carbon Dioxide. ChemSusChem 2017, 10, 1175–1185. [Google Scholar] [CrossRef]
  61. de la Cruz-Martinez, F.; Martinez, J.; Gaona, M.A.; Fernandez-Baeza, J.; Sanchez-Barba, L.F.; Rodriguez, A.M.; Castro-Osma, J.A.; Otero, A.; Lara-Sanchez, A. Bifunctional Aluminum Catalysts for the Chemical Fixation of Carbon Dioxide into Cyclic Carbonates. ACS Sustain. Chem. Eng. 2018, 6, 5322–5332. [Google Scholar] [CrossRef]
  62. Melendez, J.; North, M.; Villuendas, P. One-component catalysts for cyclic carbonate synthesis. Chem. Commun. 2009, 2577–2579. [Google Scholar] [CrossRef] [PubMed]
  63. Luo, R.; Zhou, X.; Chen, S.; Li, Y.; Zhou, L.; Ji, H. Highly efficient synthesis of cyclic carbonates from epoxides catalyzed by salen aluminum complexes with built-in “CO2 capture” capability under mild conditions. Green Chem. 2014, 16, 1496–1506. [Google Scholar] [CrossRef]
  64. Martinez, J.; de la Cruz-Martinez, F.; Gaona, M.A.; Pinilla-Penalver, E.; Fernandez-Baeza, J.; Rodriguez, A.M.; Castro-Osma, J.A.; Otero, A.; Lara-Sanchez, A. Influence of the Counterion on the Synthesis of Cyclic Carbonates Catalyzed by Bifunctional Aluminum Complexes. Inorg. Chem. 2019, 58, 3396–3408. [Google Scholar] [CrossRef]
  65. Ren, W.-M.; Liu, Y.; Lu, X.-B. Bifunctional Aluminum Catalyst for CO2 Fixation: Regioselective Ring Opening of Three-Membered Heterocyclic Compounds. J. Org. Chem. 2014, 79, 9771–9777. [Google Scholar] [CrossRef]
  66. Tian, D.; Liu, B.; Zhang, L.; Wang, X.; Zhang, W.; Han, L.; Park, D.-W. Coupling reaction of carbon dioxide and epoxides efficiently catalyzed by one-component aluminum-salen complex under solvent-free conditions. J. Ind. Eng. Chem. 2012, 18, 1332–1338. [Google Scholar] [CrossRef]
  67. Tian, D.; Liu, B.; Gan, Q.; Li, H.; Darensbourg, D.J. Formation of Cyclic Carbonates from Carbon Dioxide and Epoxides Coupling Reactions Efficiently Catalyzed by Robust, Recyclable One-Component Aluminum-Salen Complexes. ACS Catal. 2012, 2, 2029–2035. [Google Scholar] [CrossRef]
  68. Ren, Y.; Jiang, O.; Zeng, H.; Mao, Q.; Jiang, H. Lewis acid-base bifunctional aluminum-salen catalysts: Synthesis of cyclic carbonates from carbon dioxide and epoxides. RSC Adv. 2016, 6, 3243–3249. [Google Scholar] [CrossRef]
  69. 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]
  70. Hanhart, W.; Ingold, C.K. CXXXIX.—The nature of the alternating effect in carbon chains. Part XVIII. Mechanism of exhaustive methylation and its relation to anomalous hydrolysis. J. Chem. Soc. 1927, 997–1020. [Google Scholar] [CrossRef]
  71. Hughes, E.D.; Ingold, C.K.; Patel, C.S. 135. Influence of poles and polar linkings on the course pursued by elimination reactions. Part XVI. Mechanism of the thermal decomposition of quaternary ammonium compounds. J. Chem. Soc. 1933, 526–530. [Google Scholar] [CrossRef]
  72. De la Zerda, J.; Neumann, R.; Sasson, Y. Hofmann decomposition of quaternary ammonium salts under phase-transfer catalytic conditions. J. Chem. Soc. Perkin Trans. 2 1986, 823–826. [Google Scholar] [CrossRef]
  73. Hofman, A.W., IX. Contributions towards the history of the monamines.—No. III. Compound ammonias by inverse substitution. Proc. R. Soc. Lond. 1859, 10, 594–596. [Google Scholar]
  74. Collie, N.; Schryver, S.B. LIII. The action of heat on the chlorides and hydroxides of mixed quaternary ammonium compounds. J. Chem. Soc. Trans. 1890, 57, 767–782. [Google Scholar] [CrossRef] [Green Version]
  75. Zaki, A.; Fahim, H. Some quaternary ammonium salts and their decomposition products. J. Chem. Soc. 1942, 270–272. [Google Scholar] [CrossRef]
  76. Gordon, J.E. Fused Organic Salts. III.1a Chemical Stability of Molten Tetra-n-alkylammonium Salts. Medium Effects on Thermal R4N+X- Decomposition. RBr + I- = RI + Br- Equilibrium Constant in Fused Salt Medium. J. Org. Chem. 1965, 30, 2760–2763. [Google Scholar] [CrossRef]
  77. Xie, Y.; Zhang, Z.; Jiang, T.; He, J.; Han, B.; Wu, T.; Ding, K. 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]
  78. North, M.; Pasquale, R. Mechanism of Cyclic Carbonate Synthesis from Epoxides and CO2. Angew. Chem. Int. Ed. 2009, 48, 2946–2948. [Google Scholar] [CrossRef] [Green Version]
  79. Cho, W.; Shin, M.S.; Hwang, S.; Kim, H.; Kim, M.; Kim, J.G.; Kim, Y. Tertiary amines: A new class of highly efficient organocatalysts for CO2 fixations. J. Ind. Eng. Chem. 2016, 44, 210–215. [Google Scholar] [CrossRef]
Scheme 1. The syntheses of complexes 13.
Scheme 1. The syntheses of complexes 13.
Catalysts 11 00145 sch001
Scheme 2. Possible pathway for the formation of cationic species 1a and 1b.
Scheme 2. Possible pathway for the formation of cationic species 1a and 1b.
Catalysts 11 00145 sch002
Figure 1. Possible structures of cationic species of 2a and 2b found in the MS of complex 2.
Figure 1. Possible structures of cationic species of 2a and 2b found in the MS of complex 2.
Catalysts 11 00145 g001
Scheme 3. This is a figure. Schemes follow the same formatting.
Scheme 3. This is a figure. Schemes follow the same formatting.
Catalysts 11 00145 sch003
Figure 2. Kinetic curves (substrate conversion vs reaction time) of the coupling of CPO/CO2 and SO/CO2 by complex 2 under atmospheric CO2 pressure at 100 °C.
Figure 2. Kinetic curves (substrate conversion vs reaction time) of the coupling of CPO/CO2 and SO/CO2 by complex 2 under atmospheric CO2 pressure at 100 °C.
Catalysts 11 00145 g002
Figure 3. Kinetic curves of CPO/CO2 and SO/SO2 coupling catalyzed by complex 3 at 100 °C at atmospheric CO2 pressure.
Figure 3. Kinetic curves of CPO/CO2 and SO/SO2 coupling catalyzed by complex 3 at 100 °C at atmospheric CO2 pressure.
Catalysts 11 00145 g003
Table 1. The coupling reaction of terminal epoxide with CO2 by the sole complex 2 1.
Table 1. The coupling reaction of terminal epoxide with CO2 by the sole complex 2 1.
EntryEpoxideP(CO2)
(atm)
Temperature
(°C)
Time
(h)
Conversion 2
(%)
1PO1080107
2PO2080108
3PO101001068
4PO201001094
5BO1080107
6BO201001049
7CPO10801012
8CPO20801026
9CPO10100536
10CPO101001052
11CPO101002490
12CPO20100540
13CPO201001062
14SO1010055
15SO101001011
16SO101001832
17SO101002475
18SO2010057
19SO201001015
1 Conditions: Complex 2 (38.7 mg, 50 μmol), n(2)/n(epoxide) = 1/1000. 2 Determined by 1H NMR; No polymeric products were discovered as evidenced by 1H NMR.
Table 2. Complex 2 catalyzed chloropropylene oxide (CPO)/CO2 and styrene oxide (SO)/CO2 coupling under atmospheric CO2 1.
Table 2. Complex 2 catalyzed chloropropylene oxide (CPO)/CO2 and styrene oxide (SO)/CO2 coupling under atmospheric CO2 1.
Entry EpoxideTime
(h)
Conversion
(%)
1CPO56
2CPO1011
3CPO2428
4CPO3635
5CPO4868
6SO50.7
7SO102
8SO249
9SO3617
10SO4830
11SO6042
1 Conditions: Complex 2 (38.7 mg, 50 μmol), n(2)/n(epoxide) = 1/1000, P(CO2) = 1 atm, temperature is 100 °C. Other notes are the same as those in Table 1.
Table 3. The coupling reaction of terminal epoxide with CO2 by the sole complex 3 1.
Table 3. The coupling reaction of terminal epoxide with CO2 by the sole complex 3 1.
EntryEpoxideP(CO2)
(atm)
Temperature
(°C)
Time
(h)
Conversion
(%)
1PO20801014
2PO20100564
3PO201001099
4BO1080105
5BO101001076
6BO201001095
7CPO20100592
8CPO201001098
9SO201001050
10SO201002499
1 Conditions: Complex 3 (44.4 mg, 50 μmol), n(3)/n(epoxide) = 1/1000. Other notes are the same as those in Table 1.
Table 4. The coupling of epoxide/CO2 under low CO2 pressure by complex 3 1.
Table 4. The coupling of epoxide/CO2 under low CO2 pressure by complex 3 1.
EntryEpoxideP(CO2)
(atm)
Time
(h)
Conversion
(%)
1CPO2585
2CPO21098
3CPO1528
4CPO11055
5CPO12483
6CPO13691
7SO21026
8SO1510
9SO11019
10SO12450
11SO13682
12SO14892
13SO16096
1 Conditions: Complex 3 (44.4 mg, 50 μmol), n(3)/n(epoxide) = 1/1000, the temperature is 100 °C. Other notes are the same as those in Table 1.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, Z.; Wang, T.; Xiang, P.; Du, Q.; Han, S. Syntheses, Characterization, and Application of Tridentate Phenoxyimino-Phenoxy Aluminum Complexes for the Coupling of Terminal Epoxide with CO2: From Binary System to Single Component Catalyst. Catalysts 2021, 11, 145. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020145

AMA Style

Zhang Z, Wang T, Xiang P, Du Q, Han S. Syntheses, Characterization, and Application of Tridentate Phenoxyimino-Phenoxy Aluminum Complexes for the Coupling of Terminal Epoxide with CO2: From Binary System to Single Component Catalyst. Catalysts. 2021; 11(2):145. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020145

Chicago/Turabian Style

Zhang, Zhichao, Tianming Wang, Peng Xiang, Qinqin Du, and Shuang Han. 2021. "Syntheses, Characterization, and Application of Tridentate Phenoxyimino-Phenoxy Aluminum Complexes for the Coupling of Terminal Epoxide with CO2: From Binary System to Single Component Catalyst" Catalysts 11, no. 2: 145. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020145

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