Next Article in Journal
Photodegradation of Brilliant Green Dye by a Zinc bioMOF and Crystallographic Visualization of Resulting CO2
Next Article in Special Issue
Iron-Catalyzed Cross-Coupling Reactions of Alkyl Grignards with Aryl Chlorobenzenesulfonates
Previous Article in Journal
Additive Effects of Lithium Salts with Various Anionic Species in Poly (Methyl Methacrylate)
Previous Article in Special Issue
Synthesis of a Ni Complex Chelated by a [2.2]Paracyclophane-Functionalized Diimine Ligand and Its Catalytic Activity for Olefin Oligomerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Self-Assembled Bimetallic Aluminum-Salen Catalyst for the Cyclic Carbonates Synthesis

1
Department of Chemistry, Gwangju Institute of Science and Technology, Gwangju 61005, Korea
2
Department of Chemistry, University of Florida, P.O. Box 117200, Gainesville, FL 32611-7200, USA
3
SK Biotek, 325 Exporo, Yuseong-gu, Daejeon 34124, Korea
4
Gwangju Institute of Science and Technology, School of Materials Science and Engineering, Gwangju 61005, Korea
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(13), 4097; https://doi.org/10.3390/molecules26134097
Submission received: 10 May 2021 / Revised: 24 June 2021 / Accepted: 2 July 2021 / Published: 5 July 2021
(This article belongs to the Special Issue Ligands in Catalysis)

Abstract

:
Bimetallic bis-urea functionalized salen-aluminum catalysts have been developed for cyclic carbonate synthesis from epoxides and CO2. The urea moiety provides a bimetallic scaffold through hydrogen bonding, which expedites the cyclic carbonate formation reaction under mild reaction conditions. The turnover frequency (TOF) of the bis-urea salen Al catalyst is three times higher than that of a μ-oxo-bridged catalyst, and 13 times higher than that of a monomeric salen aluminum catalyst. The bimetallic reaction pathway is suggested based on urea additive studies and kinetic studies. Additionally, the X-ray crystal structure of a bis-urea salen Ni complex supports the self-assembly of the bis-urea salen metal complex through hydrogen bonding.

1. Introduction

The conversion of carbon dioxide (CO2) into value-added chemicals (carbon capture and utilization, CCU) is an area that has been increasingly studied over recent decades [1]. The conversion reaction of CO2 and epoxides to cyclic organic carbonates has a 100% atom economy, and the cyclic carbonates are important intermediates for polymers [2,3], agrochemicals, pharmaceuticals [4], electrolytes in lithium-ion batteries [5], polar aprotic solvents [6,7], and temporal protecting groups for cis-diols [8]. Owing to the usefulness of cyclic carbonate compounds, various catalysts have been developed for the reaction between CO2 and epoxides [9,10,11,12,13,14,15].
Among the developed systems, bimetallic catalysis enhances catalytic performance in cyclic carbonate synthesis reactions. In pioneering studies, Lau and Tang independently reported single-framed hetero-bimetallic (Mn and Ru) [16] and homo-bimetallic (bi-Co) [17] catalysts for the coupling reaction between CO2 and epoxides. North group also reported a μ-oxo-bridged bimetallic salen aluminum catalyst [18,19]. Since then, various bimetallic catalysts have been developed, including single-framed bimetallic catalysts [16,17,20,21,22], bridged bimetallic catalysts [18,19,23,24,25,26], ligand-separated bimetallic catalysts [27,28], and covalent bond-tethered catalysts [29,30].
In some bimetallic catalytic reactions, two metal centers can cooperatively and simultaneously activate both reaction partners (i.e., epoxide and CO2), which leads to second-order reaction kinetics [31]. For example, North group reported bimetallic aluminum catalysts, where the μ-oxo-bridged salen aluminum catalyst showed a 10-fold increase in turnover frequency (TOF) compared to a monometallic salen aluminum catalyst (Figure 1a μ-oxo-bridged catalyst) [19]. In their proposed mechanistic scenarios, one aluminum center was coordinated by ammonium-supported CO2 and the other was coordinated by the epoxide [19,32]. North and co-workers also demonstrated that the μ-oxo-bridged salen aluminum catalyst remained active in the absence of an ammonium halide cocatalyst [33]. Later, North group developed the single framed bimetallic salen aluminum catalyst (Figure 1a) [34].
We previously reported self-assembling urea-functionalized salen cobalt catalysts for the hydrokinetic resolution of epoxides [35]. Those catalysts can self-assemble in solution through weak hydrogen bonding interactions [31,35]. Herein, we report that the self-assembling bimetallic strategy can also be applied to the salen-aluminum catalysts for the cyclic carbonate synthesis from CO2 and epoxides [36,37,38,39,40,41].

2. Results and Discussion

2.1. Catalyst Preparation

Bis-urea salen ligands were synthesized in five steps following a reported procedure. The ligands were prepared from commercially available 2-tert-butylphenol, and the overall yield was 44%. All intermediate compounds were checked by 1H NMR [35]. Metalation with aluminum chloride was performed by a previously reported procedure that involved a salen ligand and di-(tert-butyl) aluminum chloride (DIBAL-Cl) (Figure 2) [42]. The formation of the bis-urea salen aluminum catalyst was confirmed by high-resolution fast atom bombardment (FAB) mass spectrometry ([M − Cl]+, m/z = calc. 1027.3356, found 1027.3505) (See Supplementary Materials for more details).

2.2. X-ray Crystallography

While we tried to grow a single crystal of a self-assembled bis-urea salen aluminum complexes, crystallization was unsuccessful. Instead, a single crystal of the bis-urea salen nickel complex was obtained by slow evaporation in N,N-dimethylformamide (DMF) (Figure 3, 4). The Ni complex (4) has same ligand as bis-urea salen catalyst (3b). As a packed crystal, the Ni complex shows a parallel head-tail conformation. In this structure, intermolecular hydrogen bonding interactions between urea groups are observed (N–H•••O = 2.06, 2.08 Å) at both ends of the salen, and the two urea planes are significantly twisted (57.9(8)°). The metal–metal distance is measured as 5.3 Å.
The dimeric structure interacts with the neighboring dimer through extended urea-urea hydrogen bonds (Figure 3). However, in this association, only the alkyl urea N-H is involved in the formation of an intermolecular N–H•••O interaction of 2.05 Å with the neighboring dimer. The aryl urea N-H forms N–H•••O interactions of 2.28 Å with one DMF molecule. Furthermore, the urea–urea plane is twisted at an angle of 53.4(1)°. The metal–metal distance between neighboring dimers is measured as 4.9 Å (Figure 4). The crystal structure of the bis-urea salen nickel complex supports the bimetallic scaffold of the bis-urea salen aluminum catalyst.

2.3. Optimization of Cyclic Carbonates Synthesis Reaciton Conditions

The initial test reaction used for this study was the conversion of propylene oxide to propylene carbonate (Figure 5). The initial reaction was conducted in a closed system consisting of a Teflon sealed Schlenk flask at 1 bar of CO2 and 45 °C. The bis-urea salen aluminum catalyst (3b) exhibited a TOF that was 13 times higher than a TOF of a standard salen-aluminum catalyst (3a) with tetrabutylammonium bromide ((n-Bu)4N+Br) (Table 1, Entries 1 and 2). Use of tetrabutylammonium iodide ((n-Bu)4N+I) led to slightly higher TOF, compared to the use of TAB under the initial reaction conditions (1 bar of CO2 and 45 °C) (Table 1, Entries 2 and 3). To increase the TOF, the reaction was conducted in a 10 mL stainless steel bomb reactor. At 10 bar of CO2 and 90 °C, the bis-urea salen aluminum catalyst showed a higher TOF than previously reported monometallic salen aluminum catalyst (Table 1, Entries 4 and 5) and bridged salen bimetallic aluminum catalyst (Table 1, Entries 5 and 6). It is important to note that the urea moiety catalyst improves the TOF regardless of the conditions of the cyclic carbonate synthesis from CO2 and epoxide reaction system.

2.4. Cyclic Carbonate Synthesis Reactions for Various Epoxides

The epoxide substrate scope was studied for the conversion reactions of various terminal epoxides (5a5h) to cyclic organic carbonates (6a6h). The bis-urea salen aluminum catalyst was most active when alkyl- or hydroxyalkyl-substituted terminal epoxides were used (5a5c, 5e, and 5g). The TOF was affected by the steric demands of the epoxides (5a5c). Chloroalkyl-substituted epoxides (5d) and styrene oxide (5f) showed low TOFs. However, alkoxy-substituted epoxides (5e, 5g, 5h) showed higher activity than other terminal epoxides, as previously reported [27]. Use of low amounts of catalyst 3b (0.02 mol%) and cocatalyst ((n-Bu)4N+I, 0.02 mol%) were allowed under comparatively mild conditions (90 °C, pCO2 = 10 bar) for various epoxide substrates (Figure 6).

2.5. Origin of Beneficial Urea Effects

In the monomeric salen aluminum catalyst system, 0.08 mol% (4 equivalent with respect to the catalyst) of urea additive was added and tested under the optimized conditions; notably, the TOF did not increase. The free urea in the reaction system did not appear to affect the reaction (Figure 7) (Table 2, Entries 1 and 2).
We speculated that the TOF enhancement could occur owing to the bimetallic character enabled by the hydrogen bonding between urea groups of the ligand. Thus, we plotted a reaction rate vs. the changes in the amount of catalyst 3b. The graph showed a clear second-order functional graph (R2 = 0.9977), suggesting a bimetallic reaction pathway (Figure 8) [35].
FTIR spectroscopy study was conducted to obtain more direct experimental evidence for self-association through urea-urea hydrogen bonding in solution. Bis-urea salen Al catalyst (3b) in THF was measured by FTIR at 25 °C. The FTIR experiments revealed the intensity of hydrogen bonded NH stretching vibration ( ν ˜ = 3444 cm−1) increased with increasing concentration, and the intensity of free NH stretching vibration ( ν ˜ = 3966 cm−1 and 3808 cm−1) decreased with increasing concentration. The results of FTIR experiments suggest intermolecular hydrogen bonding between bis-urea salen Al complexes in THF solution (Figure 9 See Supplementary Materials for more details) [35].

3. Conclusions

In conclusion, the self-assembly strategy was successfully applied to the functionalized salen-aluminum catalysts for cyclic carbonate synthesis reaction from CO2 and epoxide. The bis-urea functionalized salen ligands were designed to self-assemble through urea-urea hydrogen bonding. The bimetallic urea salen aluminum complex showed an improved reaction rate (up to 13 times at 1 bar of CO2, 45 °C and up to 2.2 times at 10 bar of CO2, 90 °C) in the cyclic carbonate formation from epoxides with CO2. Free urea additive studies and kinetic studies were performed to confirm bimetallic reaction pathway. FTIR spectroscopy study provided an experimental evidence for urea-urea hydrogen bonding in solution. This work demonstrates that hydrogen bonding can be applied to the catalysts for cyclic carbonate synthesis reaction. It is important to note that the urea moiety improves the TOF regardless of the reaction conditions. Modifications of the ligand structures to further improve the catalyst are currently in progress.

Supplementary Materials

The following are available online: 1. Materials and instrumentation; 2. Catalyst preparation; 3. Cyclic carbonates synthesis; 4. X-ray experiment; 5. Crystal structure of bis-urea salen Ni complex (4, Park4); Table S1. Crystal data and structure refinement for park4; Table S2. Atomic coordinates (× 104) and equivalent isotropic displacement parameters (Å2 × 103) for park4. U(eq) is defined as one third of the trace of the orthogonalized Uij tensor; Table S3. Bond lengths [Å] and angles [°] for park4; Table S4. Anisotropic displacement parameters (Å2 × 103) for park4. The anisotropic displacement factor exponent takes the form: −2π2[h2 a*2U11 + ... + 2 h k a* b* U12 ]; Table S5. Hydrogen coordinates (× 104) and isotropic displacement parameters (Å2 × 103) for park4; Table S6. Hydrogen bonds for park4 [Å and °].; 6. Mass spectrum of a bis-urea salen aluminum catalyst (3b); 7. FTIR spectrum of bis-urea salen aluminum catalyst (3b).

Author Contributions

Conceptualization, W.S., J.P. and S.H.; methodology, W.S. and S.H.; validation, W.S., S.K.; formal analysis, K.A.A.; investigation, S.K., W.S., H.H. and J.P.; writing—original draft preparation, H.H. and W.S.; writing—review and editing, W.S. and S.H.; supervision, S.H.; project administration, S.H.; funding acquisition, S.H. All authors contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Technology Development Program to Solve Climate Changes through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT (2017M1A2A2049102).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Baena-Moreno, F.M.; Rodríguez-Galán, M.; Vega, F.; Alonso-Fariñas, B.; Arenas, L.F.V.; Navarrete, B. Carbon capture and utilization technologies: A literature review and recent advances. Energy Sources Part A Recovery Util. Environ. Eff. 2019, 41, 1403–1433. [Google Scholar] [CrossRef]
  2. Clements, J.H. Reactive Applications of Cyclic Alkylene Carbonates. Ind. Eng. Chem. Res. 2003, 42, 663–674. [Google Scholar] [CrossRef]
  3. Paul, S.; Romain, C.; Shaw, J.; Williams, C.K. Sequence Selective Polymerization Catalysis: A New Route to ABA Block Copoly(ester-b-carbonate-b-ester). Macromolecules 2015, 48, 6047–6056. [Google Scholar] [CrossRef] [Green Version]
  4. Zhang, H.; Liu, H.-B.; Yue, J.-M. Organic Carbonates from Natural Sources. Chem. Rev. 2014, 114, 883–898. [Google Scholar] [CrossRef] [PubMed]
  5. Yun, J.; Zhang, L.; Qu, Q.; Liu, H.; Zhang, X.; Shen, M.; Zheng, H. A Binary Cyclic Carbonates-Based Electrolyte Containing Propylene Carbonate and Trifluoropropylene Carbonate for 5V Lithium-Ion batteries. Electrochim. Acta 2015, 167, 151–159. [Google Scholar] [CrossRef]
  6. Lopes, E.J.C.; Ribeiro, A.P.C.; Martins, L.M.D.R.S. New Trends in the Conversion of CO2 to Cyclic Carbonates. Catalyst 2020, 10, 479. [Google Scholar] [CrossRef]
  7. Beattie, C.; North, M.; Villuendas, P. Proline-Catalysed Amination Reactions in Cyclic Carbonate Solvents. Molecules 2011, 16, 3420–3432. [Google Scholar] [CrossRef]
  8. Laserna, V.; Fiorani, G.; Whiteoak, C.J.; Martin, E.; Escudero-Adán, E.; Kleij, A.W. Carbon dioxide as a Protecting Group: Highly Efficient and Selective Catalytic Access to Cyclic cis-Diol Scaffolds. Angew. Chem. Int. Ed. 2014, 53, 10416–10419. [Google Scholar] [CrossRef]
  9. Rehman, A.; Saleem, F.; Javed, F.; Ikhlaq, A.; Ahmad, S.W.; Harvey, A. Recent advances in the synthesis of cyclic carbonates via CO2 cycloaddition to epoxides. J. Environ. Chem. Eng. 2021, 9, 105113–105130. [Google Scholar] [CrossRef]
  10. North, M.; Pasquale, R.; Young, C. Synthesis of cyclic carbonates from epoxides and CO2. Green Chem. 2010, 12, 1514–1539. [Google Scholar] [CrossRef]
  11. 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]
  12. Rintjema, J.; Epping, R.; Fiorani, G.; Martin, E.; Escudero-Adán, E.; Kleij, A.W. Substrate-Controlled Product Divergence: Conversion of CO2 into Heterocyclic Products. Angew. Chem. Int. Ed. 2016, 55, 3972–3976. [Google Scholar] [CrossRef] [PubMed]
  13. Kim, Y.; Hyun, K.; Ahn, D.; Kim, R.; Park, M.; Kim, Y. Efficient Aluminum Catalysts for the Chemical Conversion of CO2 into Cyclic Carbonates at Room Temperature and Atmospheric CO2 Pressure. ChemSusChem 2019, 12, 4211–4220. [Google Scholar] [CrossRef]
  14. Noh, J.; Kim, Y.; Park, H.; Lee, J.; Yoon, M.; Park, M.; Kim, Y.; Kim, M. Functional group effects on a metal-organic framework catalyst for CO2 cycloaddition. J. Ind. Eng. Chem. 2018, 64, 478–483. [Google Scholar] [CrossRef]
  15. Hong, M.; Kim, Y.; Kim, H.; Cho, H.; Baik, M.; 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] [PubMed]
  16. Man, M.L.; Lam, K.C.; Sit, W.N.; Ng, S.M.; Zhou, Z.; Lin, Z.; Lau, C.P. Synthesis of Heterobimetallic Ru-Mn Complexes and the Coupling Reactions of Epoxides with Carbon Dioxide Catalyzed by these Complexes. Chem. Eur. J. 2006, 12, 1004–1015. [Google Scholar] [CrossRef]
  17. Wang, J.; Wu, J.; Tang, N. Synthesis, characterization of a new bicobalt complex [Co2L2(C2H5OH)2Cl2] and application in cyclic carbonate synthesis. Inorg. Chem. Commun. 2007, 10, 1493–1495. [Google Scholar] [CrossRef]
  18. Meléndez, 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. Clegg, W.; Harrington, R.W.; Pasquale, R.; North, M. Cyclic carbonate synthesis catalysed by bimetallic aluminium-salen complexes. Chem. Eur. J. 2010, 16, 6828–6843. [Google Scholar] [CrossRef]
  20. 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]
  21. Cozzolino, M.; Press, K.; Mazzeo, M.; Lamberti, M. Carbon dioxide/epoxide reaction catalyzed by bimetallic salalen aluminum complexes. ChemCatChem 2016, 8, 455–460. [Google Scholar] [CrossRef]
  22. Cozzolino, M.; Melchionno, F.; Santulli, F.; Mazzeo, M.; Lamberti, M. Aldimine-thioether-phenolate based mono- and bimetallic zinc complexes as catalysts for the reaction of CO2 with cyclohexene oxide. Eur. J. Inorg. Chem. 2020, 2020, 1645–1653. [Google Scholar] [CrossRef]
  23. Whiteoak, C.J.; Martin, E.; Escudero-Adán, E.; Kleij, A.W. Stereochemical Divergence in the Formation of Organic Carbonates Derived from Internal Epoxides. Adv. Synth. Catal. 2013, 355, 2233–2239. [Google Scholar] [CrossRef]
  24. Buonerba, A.; Nisi, A.D.; Grassi, A.; Milione, S.; Vagin, S.; Rieger, B.; Capacchione, C. 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]
  25. 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]
  26. Montoya, C.A.; Gómez, C.F.; Paninho, A.B.; Najdanovic-Visak, V.; Martins, L.M.D.R.S.; da Silva, M.F.C.G.; Pombeiro, A.J.L.; Ponte, M.N.; Nunes, A.V.M.; Mahmudov, K.T. Cyclic carbonate synthesis from CO2 and epoxides using zinc(II) complexes of arylhydrazones of β-diketones. J. Catal. 2016, 335, 135–140. [Google Scholar] [CrossRef] [Green Version]
  27. Whiteoak, C.J.; Martin, E.; Belmonte, M.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]
  28. Arunachalam, R.; Chinnaraja, E.; Subramanian, S.; Suresh, E.; Subramanian, P.S. Catalytic Conversion of Carbon Dioxide Using Binuclear Double-Stranded Helicates: Cyclic Carbonate from Epoxides and Diol. ACS Omega 2020, 5, 14890–14899. [Google Scholar] [CrossRef]
  29. Maeda, C.; Taniguchi, T.; Ogawa, K.; Ema, T. Bifunctional Catalysts Based on m-Phenylene-Bridged Porphyrin Dimer and Trimer Platforms: Synthesis of Cyclic Carbonates from Carbon Dioxide and Epoxides. Angew. Chem. Int. Ed. 2015, 54, 134–138. [Google Scholar] [CrossRef]
  30. He, S.; Wang, F.; Tong, W.-L.; Yiu, S.-M.; Chan, M.C.W. Topologically diverse shape-persistent bis-(Zn–salphen) catalysts: Efficient cyclic carbonate formation under mild conditions. Chem. Commun. 2016, 52, 1017–1020. [Google Scholar] [CrossRef]
  31. Park, J.; Hong, S. Cooperative bimetallic catalysis in asymmetric transformations. Chem. Soc. Rev. 2012, 41, 6931–6943. [Google Scholar] [CrossRef]
  32. Beattie, C.; Villuendas, P.; Young, C.; North, M. Influence of Temperature and Pressure on Cyclic Carbonate Synthesis Catalyzed by Bimetallic Aluminum Complexes and Application to Overall syn-Bis-hydroxylation of Alkenes. J. Org. Chem. 2013, 78, 419–426. [Google Scholar] [CrossRef]
  33. Castro-Osma, J.A.; Leitner, W.; Muller, T.E.; North, M.; Offermans, W.K. Unprecedented Carbonato Intermediates in Cyclic Carbonate Synthesis Catalysed by Bimetallic Aluminium(Salen) Complexes. ChemSusChem 2016, 9, 791–794. [Google Scholar] [CrossRef]
  34. 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]
  35. Park, J.; Lang, K.; Abboud, K.A.; Hong, S. Self-Assembly Approach toward Chiral Bimetallic Catalysts: Bis-Urea-Functionalized (Salen)Cobalt Complexes for the Hydrolytic Kinetic Resolution of Epoxides. Chem. Eur. J. 2011, 17, 2236–2245. [Google Scholar] [CrossRef] [PubMed]
  36. Broghammer, F.; Brodbeck, D.; Junge, T.; Peters, R. Cooperative Lewis acid−onium salt catalysis as tool for the desymmetrization of meso-epoxides. Chem. Commun. 2017, 53, 1156–1159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Deacy, A.C.; Moreby, E.; Phanopoulos, A.; Williams, C.K. Co(III)/Alkali-Metal(I) Heterodinuclear Catalysts for the Ring-Opening Copolymerization of CO2 and Propylene Oxide. J. Am. Chem. Soc. 2020, 142, 19150–19160. [Google Scholar] [CrossRef] [PubMed]
  38. Deacy, A.C.; Durr, C.B.; Kerr, R.W.F.; Williams, C.K. Heterodinuclear catalysts Zn(II)/M and Mg(II)/M, where M = Na(I), Ca(II) or Cd(II), for phthalic anhydride/cyclohexene oxide ring opening copolymerisation. Catal. Sci. Technol. 2021, 11, 3109–3118. [Google Scholar] [CrossRef]
  39. Noh, E.; Na, S.; Sujith, S.; Kim, S.-W.; Lee, B. Two Components in a Molecule: Highly Efficient and Thermally Robust Catalytic System for CO2/Epoxide Copolymerization. J. Am. Chem. Soc. 2007, 129, 8082–8083. [Google Scholar] [CrossRef]
  40. Lidston, C.A.L.; Abel, B.A.; Coates, G.W. Bifunctional Catalysis Prevents Inhibition in Reversible-Deactivation Ring-Opening Copolymerizations of Epoxides and Cyclic Anhydrides. J. Am. Chem. Soc. 2020, 142, 20161–20169. [Google Scholar] [CrossRef]
  41. Hubbell, A.K.; Coates, G.W. Nucleophilic Transformations of Lewis Acid-Activated Disubstituted Epoxides with Catalyst-Controlled Regioselectivity. J. Org. Chem. 2020, 85, 13391–13414. [Google Scholar] [CrossRef] [PubMed]
  42. Quek, S.C.Z.; Pridmore, N.E.; Whitwood, A.C.; Wu, X.; North, M. Aluminum(salen) Complexes as Catalysts for the Kinetic Resolution of Terminal Epoxides via CO2 Coupling. ACS Catal. 2015, 5, 3398–3402. [Google Scholar]
Figure 1. Bimetallic catalysts for cyclic carbonates synthesis. (a) Reported bimetallic salen catalysts. (b) This research: self-assembled bimetallic catalysts through hydrogen bonding.
Figure 1. Bimetallic catalysts for cyclic carbonates synthesis. (a) Reported bimetallic salen catalysts. (b) This research: self-assembled bimetallic catalysts through hydrogen bonding.
Molecules 26 04097 g001
Figure 2. Synthesis of salen-aluminum catalysts.
Figure 2. Synthesis of salen-aluminum catalysts.
Molecules 26 04097 g002
Figure 3. X-ray crystal structure of bis-urea salen nickel complex 4.
Figure 3. X-ray crystal structure of bis-urea salen nickel complex 4.
Molecules 26 04097 g003
Figure 4. Hydrogen-bond packing structure of bis-urea salen nickel complex.
Figure 4. Hydrogen-bond packing structure of bis-urea salen nickel complex.
Molecules 26 04097 g004
Figure 5. Cyclic carbonate synthesis using salen-Al catalysts and tetrabutylammonium halides.
Figure 5. Cyclic carbonate synthesis using salen-Al catalysts and tetrabutylammonium halides.
Molecules 26 04097 g005
Figure 6. Scopes of epoxides. Reaction conditions: Epoxide (47 mmol), CO2 (10 bar), Catalyst (10 mg, 0.0094 mmol), (n-Bu)4N+I (3.5 mg, 0.0094 mmol), reaction was carried out in a 10 mL sealed stainless steel pressure bomb reactor. All TOFs are determined by 1H NMR using mesitylene internal standard (1.0 mmol, 0.1 mL). 1 3.5 h reaction.
Figure 6. Scopes of epoxides. Reaction conditions: Epoxide (47 mmol), CO2 (10 bar), Catalyst (10 mg, 0.0094 mmol), (n-Bu)4N+I (3.5 mg, 0.0094 mmol), reaction was carried out in a 10 mL sealed stainless steel pressure bomb reactor. All TOFs are determined by 1H NMR using mesitylene internal standard (1.0 mmol, 0.1 mL). 1 3.5 h reaction.
Molecules 26 04097 g006
Figure 7. Urea additive study.
Figure 7. Urea additive study.
Molecules 26 04097 g007
Figure 8. Kinetic analysis of the reaction order (Second order, R2 = 0.9977). Reaction conditions: Epoxide (47 mmol), CO2 (10 bar), (n-Bu)4N+I (3.5 mg, 0.0094 mmol), 90 °C, 3 h, reaction was carried out in a 10 mL sealed stainless steel pressure bomb reactor. All TOFs are determined by 1H NMR using mesitylene internal standard (1.0 mmol, 0.1 mL).
Figure 8. Kinetic analysis of the reaction order (Second order, R2 = 0.9977). Reaction conditions: Epoxide (47 mmol), CO2 (10 bar), (n-Bu)4N+I (3.5 mg, 0.0094 mmol), 90 °C, 3 h, reaction was carried out in a 10 mL sealed stainless steel pressure bomb reactor. All TOFs are determined by 1H NMR using mesitylene internal standard (1.0 mmol, 0.1 mL).
Molecules 26 04097 g008
Figure 9. The NH stretching region of the FTIR spectra of 3b in THF at two different concentration (1 mM (grey), 1.5 mM (black)) at 25 °C.
Figure 9. The NH stretching region of the FTIR spectra of 3b in THF at two different concentration (1 mM (grey), 1.5 mM (black)) at 25 °C.
Molecules 26 04097 g009
Table 1. Cyclic carbonate synthesis using salen-Al catalysts and tetrabutylammonium halides.
Table 1. Cyclic carbonate synthesis using salen-Al catalysts and tetrabutylammonium halides.
EntryCatalyst (mol%)Cocatalyst (mol%)CO2 (Bar)T (°C)TOF (h−1) 1
1 23a (1)(n-Bu)4N+Br (1)1451
23b (0.02)(n-Bu)4N+Br (0.02)14513
33b (0.02)(n-Bu)4N+I (0.02)14517
43a (0.02)(n-Bu)4N+I (0.02)109065
53b (0.02)(n-Bu)4N+I (0.02)1090144
61 (0.02)(n-Bu)4N+I (0.02)109047 3
1 Determined by 1H NMR using mesitylene internal standard 2 12 h reaction 3 TOF = TON/h·[Al].
Table 2. Urea additive study.
Table 2. Urea additive study.
EntryUrea Additive (mol%)TOF (h−1)
1-64
20.08 mol%65
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Seong, W.; Hahm, H.; Kim, S.; Park, J.; Abboud, K.A.; Hong, S. Self-Assembled Bimetallic Aluminum-Salen Catalyst for the Cyclic Carbonates Synthesis. Molecules 2021, 26, 4097. https://doi.org/10.3390/molecules26134097

AMA Style

Seong W, Hahm H, Kim S, Park J, Abboud KA, Hong S. Self-Assembled Bimetallic Aluminum-Salen Catalyst for the Cyclic Carbonates Synthesis. Molecules. 2021; 26(13):4097. https://doi.org/10.3390/molecules26134097

Chicago/Turabian Style

Seong, Wooyong, Hyungwoo Hahm, Seyong Kim, Jongwoo Park, Khalil A. Abboud, and Sukwon Hong. 2021. "Self-Assembled Bimetallic Aluminum-Salen Catalyst for the Cyclic Carbonates Synthesis" Molecules 26, no. 13: 4097. https://doi.org/10.3390/molecules26134097

Article Metrics

Back to TopTop