Next Article in Journal
De-Escalation of Saccharification Costs through Enforcement of Immobilization of Cellulase Synthesized by Wild Trichoderma viride
Next Article in Special Issue
Efficient Removal of Organic Dye from Aqueous Solution Using Hierarchical Zeolite-Based Biomembrane: Isotherm, Kinetics, Thermodynamics and Recycling Studies
Previous Article in Journal
Recent Developments in Nanocatalyzed Green Synthetic Protocols of Biologically Potent Diverse O-Heterocycles—A Review
Previous Article in Special Issue
Partial Hydrogenation of Soybean and Waste Cooking Oil Biodiesel over Recyclable-Polymer-Supported Pd and Ni Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Iridium(triNHC)-Catalyzed Transfer Hydrogenation of Glycerol Carbonate without Exogenous Reductants

Department of Energy Systems Research, Ajou University, Suwon 16499, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(6), 656; https://doi.org/10.3390/catal12060656
Submission received: 25 April 2022 / Revised: 7 June 2022 / Accepted: 13 June 2022 / Published: 15 June 2022
(This article belongs to the Special Issue Exclusive Papers of the Editorial Board Members (EBMs) of Catalysts)

Abstract

:
The iridium(Ir) (triNHC = tri-N-heterocyclic carbene)-catalyzed transfer hydrogenation of glycerol carbonate (GC) is described in the absence of additional hydride sources. The described reduction provides a sustainable route to produce industrially-valuable formate and lactate with high turnover numbers (TONs). The bimetallic Ir(I) involving triNHC carbene ligands exhibits high TONs, and the reaction mechanism, including the bimetallic Ir(triNHC) catalyst, is proposed based on mechanistic studies.

Graphical Abstract

1. Introduction

An indirect reduction in carbon dioxide (CO2) via hydrogenation and transfer hydrogenation of organic carbonates is a practical and eco-friendly strategy that addresses global warming, plastic pollution, and energy problems by reducing the concentration of CO2 in the air, recycling polycarbonate-based plastics, and producing valuable fuels and chemical feedstocks (formic acid, methanol, and diols) [1,2,3,4,5]. Thus, various transition metal-catalyzed reductions in organic carbonates have been studied, although organic carbonates are not highly susceptible to reduction. Ruthenium catalysts modified with tridentate ligands such as phosphorous-nitrogen-nitrogen (PNN) [6,7], phosphorous-nitrogen-phosphorous(PNP) [8,9], and carbon-nitrogen-carbon (CNC) [10], manganese catalysts involving PNN [11,12] and PNP [13] ligands, and cobalt catalysts containing nitrogen-oxygen coordinating ligands [14] are used to hydrogenate organic carbonates.
Although the transition metal-catalyzed hydrogenation of carbonates exhibits acceptable yields and catalytic activities, high pressure of H2 (30–60 bar) should be applied, and the reactions must be run in the pressurized equipment. Thus, the isopropanol (IPA)-mediated reduction in organic carbonates, called transfer hydrogenation, is investigated. Ru(PNP) [15], Fe(PNP) [16], and Co(NO) [14] complexes have been used to reduce organic carbonates. In addition to transition metal-catalyzed hydrogenation and transfer hydrogenation, boron- and silane-mediated reduction in carbonates were reported in the presence of MgBu2 [17], halide anions [18], and B(C6F5)3 [19] catalysts.
For the transfer hydrogenation of organic carbonates, alcohols such as IPA and glycerol are required. IPA is a hydrogen source delivering hydrides to carbonates, generating reduced products (formic acid) and byproducts (acetone) formed from IPA [20]. Biomass-derived glycerol has been recently used for transfer hydrogenation with the environmental advantages of sustainable hydrogen sources. However, glycerol has not been employed to reduce organic carbonates [21,22]. Given that our research group has been studying glycerol-mediated transfer hydrogenation [23,24,25], the reaction of glycerol carbonate (GC) derived from CO2 and glycerol has drawn our interest [26,27], where the additional reductants may not be required due to the glycerol moiety in GC.
As illustrated in Scheme 1, GC involves both C1 and C3 sources in the molecule. With a judicious choice of catalysts, GC plays C1 and C3 sources along with hydride sources. The transfer hydrogenation of GC produces formate (C1 product) and lactate (C3 product), considered essential sustainable energy sources and bioplastic raw materials, respectively [28,29,30,31]. In this study, we report the first Ir(triNHC)-catalyzed transfer hydrogenation of GC in the absence of exogenous hydrides. The reaction mechanism is probed based on control experiments, and the origin of the high catalytic activity of bimetallic Ir(triNHC) was speculated based on the proposed reaction mechanism.

2. Results and Discussion

The reaction optimization results of transfer hydrogenation of GC are presented in Table 1. The reactions of GC, bases, and iridium catalysts were conducted in N-methyl-2-pyrrolidone (NMP) at 200 °C for 20 h. Catalysts AD are reported [23,32,33,34], and catalysts E and F are newly synthesized. The optimization begins with different loadings of catalyst A and CsOH⋅H2O (2 equiv.), showing turnover numbers (TONs) of 6500 (formate) and 9600 (lactate) with catalyst A (2.5 × 10−3 mol%) (entries 1–3). The TONs of formate and lactate were calculated by 1H NMR spectroscopy in D2O using isonicotinic acid as an internal standard (see Supplementary Materials Figure S5). Changing the solvent to H2O diminished the catalytic activity (entry 4). Both NMP and water are commonly used solvents in the reactions of glycerol. Depending on the catalytic system, the optimized solvent was varied, and our catalytic system exhibited higher TONs in NMP. Although NMP can be converted to 4-N-methylaminobutanoate in the presence of bases, it has been known that 4-N-methylaminobutanoate does not affect the catalytic reaction of glycerol [35]. Different bases (KOH and NaOH) exhibit inferior results in terms of TONs of each product (entries 5 and 6). Increasing the amounts of CsOH⋅H2O did not improve TONs (entry 7). In the absence of bases or catalysts, no product was observed (entries 8 and 9). The monometallic Ir catalysts BD exhibit slightly lower TONs compared to bimetallic catalyst A (entry 10–13). The catalyst loadings of monometallic catalysts BD were determined to maintain the same mole numbers of iridium ions of bimetallic catalyst A. Analogous to catalyst A, monometallic catalyst B showed increased TONs with lower catalyst loadings (entries 10 and 11). Ir(III) catalysts E and F involving the carbene, pyridine, and pentamethylcyclopentadiene (Cp*) show much lower TONs compared to Ir(I) catalysts AD modified with carbene ligands (entries 14 and 15). The reactions of ethylene carbonate and propylene carbonate formed formate with TONs of 580 and 415, respectively, under the conditions of entry 2 (Table 1) (see Supplementary Materials Figures S6 and S7). Thus, the presence of a hydroxy group in GC is critical for high TONs. The stability of Ir(triNHC) (catalyst A) was evaluated by accumulation experiments, showing a slight decrease in TONs of each product in the second reaction (see Supplementary Materials Scheme S3).
The mixture of catalysts, glycerol carbonate (10.6 mmol), and base in NMP (10 mL) was heated at 200 °C for 20 h. Catalysts A and E have two iridium ions in the molecule.
Catalysts 12 00656 i002
Control experiments were conducted to examine the transfer hydrogenation mechanism of GC (Scheme 2 and Scheme 3). Because GC might be dissociated to form glycerol and carbonates (CO32-) in the presence of CsOH, the reactions of Cs2CO3 with glycerol were performed with catalyst A (Scheme 2). Compared to TONs of the GC reaction (FA = 3600, LA = 5300, entry 2 of Table 1), the reaction of Cs2CO3 with glycerol produces products with slightly lower TONs (FA = 3000 and LA = 4300). Thus, the reaction mechanism may involve the dissociation of GC into carbonate anions and glycerol.
The reactions of GC with additional alcohols (IPA and glycerol) were implemented with the assumption that an additional hydrogen source may increase the yields of formate and lactate. As depicted in Equation (1) of Scheme 3, the addition of IPA reduced the TONs of each product, and additional glycerol also slightly reduced the TONs of products. Because the additional alcohols reacted with catalysts and OH- to induce dehydrogenation, glycerol carbonate reacted with reduced amounts of catalysts and bases, resulting in lower TONs of formate and lactate. An isotope labeling experiment using deuterated glycerol was conducted (Equation (2) of Scheme 3), forming small amounts of deuterated formate (approximately less than 10% by mass spectrometry analysis, see Supplementary Materials Figure S8). Accordingly, most hydrogen atoms in the formate were derived from GC, and approximately 10% of external hydrogens were incorporated in the formate.
The gas analysis of the reaction mixture was conducted to monitor the generation of CO2 and H2 from GC (Figure 1). Briefly, gaseous products were sampled from the headspace of the reactor at reaction times of 3 h and 20 h and then analyzed using gas chromatography. The number of moles of each product was determined from the area of the corresponding peak in the measured chromatogram. As depicted in Figure 1, H2 gas (0.06 mmol) is generated at 3 h and 0.58 mmol at 20 h. The generated CO2 was very low (as illustrated in the inset graph), where 0.53 and 0.31 µmol of CO2 were observed at 3 and 20 h, respectively. During the reaction of GC, simple dehydrogenation forming H2 gas is not a significant route based on the amount of generated H2 (0.58 mmol) and used GC (10.6 mmol). The dissociation of CO2 from GC rarely occurs based on the amount of liberated CO2. We exclude the hydrogenation of CO2 as the main route of the mechanism because significant amounts of H2 and CO2 were not generated for the hydrogenation mechanism based on control experiments and gas analysis.
We proposed a catalytic cycle of Ir(triNHC)-catalyzed transfer hydrogenation of GC (Scheme 4). Catalyst A releases cyclooctadiene (COD) and coordinates to the glycerol alkoxide and bicarbonate [23]. Glycerol alkoxides and bicarbonates are formed from GC and OH based on 13C NMR of the mixture of GC and OH (see Supplementary Materials Figure S9). Intermediate I undergoes β-hydrogen elimination to produce intermediate II. Dihydroxyacetone (DHA) dissociated from intermediate II is converted to lactate via dehydration and Cannizzaro reaction [36]. The elimination of CsOH from intermediate III affords CO2-bound intermediate IV. The co-operative interaction of two iridium ions in intermediate III promotes the elimination of CsOH via intermediates III’ and III’’ [37,38]. The hydrometallation of CO2 produces formic acid. GC and CsOH enter the catalytic cycle after releasing formates. The bimetallic catalysts showed higher TONs than monometallic catalysts; catalyst A vs. catalysts BD. The efficient reaction of Ir-H with CO2 might be promoted by the co-operative action of two iridium ions in the bimetallic catalyst’s reaction sphere.

3. Conclusions

We report the first transfer hydrogenation of GC without additional hydrogen sources to the best of our knowledge. The control experiments and gas analysis confirm the reduction route induced by Ir(triNHC) catalysts. The reaction results of GC with deuterated glycerol imply that dissociated bicarbonates and glycerol alkoxides from GC are readily reacted before the participation of the external hydride sources in the reaction. The bimetallic catalysts are favored in coordinating both reactants (bicarbonates and glycerol alkoxides) inside the reaction sphere, as illustrated in the proposed reaction mechanism.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12060656/s1, Scheme S1. The reaction of ethylene carbonate, Scheme S2. The reaction of propylene carbonate, Scheme S3. Accumulation experiments of GC and OH, Table S1. Transfer hydrogenation of glycerol carbonate, Figure S1. 1H NMR spectra of catalyst E, Figure S2. 13C NMR spectra of catalyst E, Figure S3. 1H NMR spectra of catalyst F, Figure S4. 13C NMR spectra of catalyst F, Figure S5. 1H NMR spectra of Entry 2 (Table 1) indicating the signature peak of each species, Figure S6. 1H NMR spectra of the reaction mixture using ethylene glycol, Figure S7. 1H NMR spectra of the reaction mixture using propylene glycol, Figure S8. Mass spectrum of the reaction mixture involving d-glycerol, Figure S9. 13C NMR spectra; glycerol (green), CsHCO3 (red), the mixture of glycerol carbonate (GC) and CsOH (blue) and GC (black), Figure S10. 1H NMR spectra in DMSO-d6; (a) glycerol carbonate (green), (b) glycerol (blue), (c) catalyst A (red), and (d) the mixture of glycerol carbonate (1 equiv), CsOH (2 equiv), catalyst A (0.5 equiv) (black). References [23,33,34] have been in the Supplementary Materials.

Author Contributions

Conceptualization, H.-Y.J.; Investigations, Y.-J.C., M.-h.L., H.B., J.P. and S.Y.; writing—original draft preparation, H.-Y.J.; writing—review and editing, H.-Y.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Research Foundation of Korea, grant number 2020M3H7A1098283 and 2022R1A2C1004387.

Data Availability Statement

The data presented in this study are openly available in the Supplementary Materials.

Acknowledgments

This study was supported by the Carbon to X Program (No. 2020M3H7A1098283) and Basic Science Research Program (No. 2022R1A2C1004387) by the National Research Foundation of Korea, which is funded by the Ministry of Science and ICT).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sordakis, K.; Tang, C.; Vogt, L.K.; Junge, H.; Dyson, P.J.; Beller, M.; Laurenczy, G. Homogeneous catalysis for sustainable hydrogen storage in formic acid and alcohols. Chem. Rev. 2018, 118, 372–433. [Google Scholar] [CrossRef] [PubMed]
  2. Jessop, P.G.; Ikariya, T.; Noyori, R. Homogeneous hydrogenation of carbon-dioxide. Chem. Rev. 1995, 95, 259–272. [Google Scholar] [CrossRef]
  3. Jessop, P.G.; Joó, F.; Tai, C.C. Recent advances in the homogeneous hydrogenation of carbon dioxide. Coord. Chem. Rev. 2004, 248, 2425–2442. [Google Scholar] [CrossRef]
  4. Wang, W.-H.; Himeda, Y.; Muckerman, J.T.; Manbeck, G.F.; Fujita, E. CO2 Hydrogenation to formate and methanol as an alternative to photo- and electrochemical CO2 reduction. Chem. Rev. 2015, 115, 12936–12973. [Google Scholar] [CrossRef]
  5. Dabral, S.; Schaub, T. The use of carbon dioxide (CO2) as a building block in organic synthesis from an industrial perspective. Adv. Synth. Catal. 2019, 361, 223–246. [Google Scholar] [CrossRef] [Green Version]
  6. Balaraman, E.; Gunanathan, C.; Zhang, J.; Shimon, L.J.W.; Milstein, D. Efficient hydrogenation of organic carbonates, carbamates and formates indicates alternative routes to methanol based on CO2 and CO. Nat. Chem. 2011, 3, 609–614. [Google Scholar] [CrossRef]
  7. Krall, E.M.; Klein, T.W.; Andersen, R.J.; Nett, A.J.; Glasgow, R.W.; Reader, D.S.; Dauphinais, B.C.; Mc Ilrath, S.P.; Fischer, A.A.; Carney, M.J.; et al. Controlled hydrogenative depolymerization of polyesters and polycarbonates catalyzed by ruthenium(II) PNN pincer complexes. Chem. Commun. 2014, 50, 4884–4887. [Google Scholar] [CrossRef]
  8. Han, Z.; Rong, L.; Wu, J.; Zhang, L.; Wang, Z.; Ding, K. Catalytic hydrogenation of cyclic carbonates: A practical approach from CO2 and epoxides to methanol and diols. Angew. Chem. Int. Ed. 2012, 51, 13041–13045. [Google Scholar] [CrossRef]
  9. Li, Y.; Junge, K.; Beller, M. Improving the efficiency of the hydrogenation of carbonates and carbon dioxide to methanol. ChemCatChem 2013, 5, 1072–1074. [Google Scholar] [CrossRef]
  10. Wu, X.; Ji, L.; Ji, Y.; Elageed, E.H.M.; Gao, G. Hydrogenation of ethylene carbonate catalyzed by lutidine-bridged N-heterocyclic carbene ligands and ruthenium precursors. Catal. Commun. 2016, 85, 57–60. [Google Scholar] [CrossRef]
  11. Kumar, A.; Janes, T.; Espinosa-Jalapa, N.A.; Milstein, D. Manganese catalyzed hydrogenation of organic carbonates to methanol and alcohols. Angew. Chem. Int. Ed. 2018, 57, 12076–12080. [Google Scholar] [CrossRef] [PubMed]
  12. Zubar, V.; Levedev, Y.; Azofra, L.M.; Cavallo, L.; El-Sepelgy, O.; Rueping, M. Hydrogenation of CO2-derived carbonates and polycarbonates to methanol and diols by metal-ligand cooperative manganese catalysis. Angew. Chem. Int. Ed. 2018, 57, 13439–13443. [Google Scholar] [CrossRef] [PubMed]
  13. Kaithal, A.; Hölscher, M.; Leitner, W. Catalytic hydrogenation of cyclic carbonates using manganese complexes. Angew. Chem. Int. Ed. 2018, 57, 13449–13453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Dahiya, P.; Gangwar, M.K.; Sundararaju, B. Well-defined Cp*Co(III)-catalyzed hydrogenation of carbonates and polycarbonates. ChemCatChem 2021, 13, 934–939. [Google Scholar] [CrossRef]
  15. Kim, S.H.; Hong, S.H. Transfer hydrogenation of organic formates and cyclic carbonates: An alternative route to methanol from carbon dioxide. ACS Catal. 2014, 4, 3630–3636. [Google Scholar] [CrossRef]
  16. Liu, X.; de Vries, J.G.; Werner, T. Transfer hydrogenation of cyclic carbonates and polycarbonate to methanol and diols by iron pincer catalysts. Green Chem. 2019, 21, 5248–5255. [Google Scholar] [CrossRef]
  17. Szewczyk, M.; Magre, M.; Zubar, V.; Rueping, M. Reduction of cyclic and linear organic carbonates using a readily available magnesium catalyst. ACS Catal. 2019, 9, 11634–11639. [Google Scholar] [CrossRef]
  18. Bobbink, F.D.; Menoud, F.; Dyson, P.J. Synthesis of methanol and diols from CO2 via cyclic carbonates under metal-free, ambient pressure, and solvent-free conditions. ACS Sustain. Chem. Eng. 2018, 6, 12119–12123. [Google Scholar] [CrossRef]
  19. Feghali, E.; Cantat, T. Room temperature organocatalyzed reductive depolymerization of waste polyethers, polyesters, and polycarbonates. ChemSusChem 2015, 8, 980–984. [Google Scholar] [CrossRef]
  20. Zassinovich, G.; Mestroni, G.; Gladiali, S. Asymmetric hydrogen transfer reactions promoted by homogeneous transition metal catalysis. Chem. Rev. 1992, 92, 1051–1069. [Google Scholar] [CrossRef]
  21. Díaz-Álvarez, A.E.; Cadierno, V. Glycerol: A promising green solvent and reducing agent for metal-catalyzed transfer hydrogenation reactions and nanoparticles formation. Appl. Sci. 2013, 3, 55–69. [Google Scholar] [CrossRef]
  22. Crabtree, R.H. Transfer hydrogenation with glycerol as H-donor: Catalyst activation, deactivation and homogeneity. ACS Sustain. Chem. Eng. 2019, 7, 15845–15853. [Google Scholar] [CrossRef]
  23. Cheong, Y.-J.; Sung, K.; Park, S.; Jung, J.; Jang, H.-Y. Valorization of chemical wates: Ir(biscarbene)-catalyzed transfer hydrogenation of inorganic carbonates using glycerol. ACS Sustain. Chem. Eng. 2020, 8, 6972–6978. [Google Scholar] [CrossRef]
  24. Sung, K.; Lee, M.-h.; Cheong, Y.-J.; Jang, H.-Y. Ir(triscarbene)-catalyzed sustainable transfer hydrogenation of levulinic acid to γ-valerolactone. Appl. Organomet. Chem. 2020, 35, e6105. [Google Scholar] [CrossRef]
  25. Cheong, Y.-J.; Sung, K.; Kim, J.-A.; Kim, Y.K.; Yoon, W.; Yun, H.; Jang, H.-Y. Iridium(NHC)-catalyzed sustainable transfer hydrogenation of CO2 and inorganic carbonates. Catalysts 2021, 11, 695. [Google Scholar] [CrossRef]
  26. Christy, S.; Noschese, A.; Lomelí-Rodrigeuz, M.; Greeves, N.; Lopez-Sanchez, J.A. Recent progress in the synthesis and applications of glycerol carbonate. Curr. Opin. Green Sustain. Chem. 2018, 14, 99–107. [Google Scholar] [CrossRef]
  27. Szöri, M.; Giri, B.R.; Wang, Z.; Dawood, A.E.; Viskolcz, B.; Farooq, A. Glycerol carbonate as a fuel additive for a sustainable future. Sustain. Energy Fuels 2018, 2, 2171–2178. [Google Scholar] [CrossRef] [Green Version]
  28. Loges, B.; Boddien, A.; Gartner, F.; Junge, H.; Beller, M. Catalytic generation of hydrogen from formic acid and its derivatives: Useful hydrogen storage materials. Top. Catal. 2010, 53, 902–914. [Google Scholar] [CrossRef]
  29. Grasemann, M.; Laurenczy, G. Formic acid as a hydrogen source-recent developments and future trends. Energy Environ. Sci. 2012, 5, 8171–8181. [Google Scholar] [CrossRef]
  30. Mellmann, D.; Sponholz, P.; Junge, H.; Beller, M. Formic acid as a hydrogen storage material-development of homogeneous catalysts for selective hydrogen release. Chem. Soc. Rev. 2016, 45, 3954–3988. [Google Scholar] [CrossRef]
  31. Bottari, G.; Barta, K. Lactic acid and hydrogen from glycerol via acceptorless dehydrogenation using homogeneous catalysts. Recycl. Catal. 2015, 2, 70–77. [Google Scholar] [CrossRef] [Green Version]
  32. Frey, G.D.; Rentzsch, C.F.; von Preysing, D.; Scherg, T.; Mühlhofer, M.; Herdtweck, E.; Herrmann, W.A. Rhodium and iridium complexes of N-heterocyclic carbenes: Structural investigations and their catalytic properties in the borylation reaction. J. Organomet. Chem. 2006, 691, 5725–5738. [Google Scholar] [CrossRef]
  33. Sharninghausen, L.S.; Campos, J.; Manas, M.G.; Crabtree, R.H. Efficient selective and atom economic catalytic conversion of glycerol to lactic acid. Nat. Commun. 2014, 5, 5084. [Google Scholar] [CrossRef] [PubMed]
  34. Cheong, Y.-J.; Sung, K.; Kim, J.-A.; Kim, Y.K.; Jang, H.-Y. Highly efficient iridium-catalyzed production of hydrogen and lactate from glycerol: Rapid hydrogen evolution by bimetallic iridium catalysts. Eur. J. Inorg. Chem. 2020, 2020, 4064–4068. [Google Scholar] [CrossRef]
  35. Sharninghausen, L.S.; Mercado, B.Q.; Crabtree, R.H.; Hazari, N. Selective conversionof glycerol to lactic acid with iron pincer precatalysts. Chem. Commun. 2015, 51, 16201–16204. [Google Scholar] [CrossRef]
  36. Lu, Z.; Cherepakhin, V.; Demianets, I.; Lauridsen, P.J.; Williams, T.J. Iridium-based hydride transfer catalysts: From hydrogen storage to fine chemicals. Chem. Commun. 2018, 54, 7711–7724. [Google Scholar] [CrossRef]
  37. Truscott, B.J.; Kruger, H.; Webb, P.B.; Bühl, M.; Nolan, S.P. The mechanism of CO2 insertion into iridium(I) hydroxide and alkoxide bonds: A kinetics and computational study. Chem. Eur. J. 2015, 21, 6930–6935. [Google Scholar] [CrossRef]
  38. Vummaleti, S.V.C.; Talarico, G.; Nolan, S.P.; Cavallo, L.; Poater, A. Mechanism of CO2 fixation by IrI-X bonds (X = OH, OR, N, C). Eur. J. Inorg. Chem. 2015, 2015, 4653–4657. [Google Scholar] [CrossRef]
Scheme 1. Transfer hydrogenation of GC.
Scheme 1. Transfer hydrogenation of GC.
Catalysts 12 00656 sch001
Scheme 2. Reaction of Cs2CO3 and glycerol.
Scheme 2. Reaction of Cs2CO3 and glycerol.
Catalysts 12 00656 sch002
Scheme 3. Reaction of GC in the presence of additional alcohols.
Scheme 3. Reaction of GC in the presence of additional alcohols.
Catalysts 12 00656 sch003
Figure 1. Gas analysis from the reaction mixture of glycerol carbonate, catalyst A, and CsOH.
Figure 1. Gas analysis from the reaction mixture of glycerol carbonate, catalyst A, and CsOH.
Catalysts 12 00656 g001
Scheme 4. Plausible reaction mechanism.
Scheme 4. Plausible reaction mechanism.
Catalysts 12 00656 sch004
Table 1. Transfer hydrogenation of GC.
Table 1. Transfer hydrogenation of GC.
Catalysts 12 00656 i001
EntryCatalyst
(mol%)
Base
(Equiv.)
SolventFormate
(TON)
Lactate
(TON)
1A (2.5 × 10−3)CsOH⋅H2O (2)NMP65009600
2A (5.0 × 10−3)CsOH⋅H2O (2)NMP36005300
3A (1.0 × 10−2)CsOH⋅H2O (2)NMP9801900
4A (5.0 × 10−3)CsOH⋅H2O (2)H2O2189
5A (5.0 × 10−3)KOH (2)NMP1265
6A (5.0 × 10−3)NaOH (2)NMP160130
7A (5.0 × 10−3)CsOH⋅H2O (3)NMP9304800
8A (5.0 × 10−3)--NMP----
9--CsOH⋅H2O (2)NMP----
10B (1.0 × 10−2)CsOH⋅H2O (2)NMP24003600
11B (5.0 × 10−3)CsOH⋅H2O (2)NMP52007400
12C (1.0 × 10−2)CsOH⋅H2O (2)NMP26004200
13D (1.0 × 10−2)CsOH⋅H2O (2)NMP21003400
14E (5.0 × 10−3)CsOH⋅H2O (2)NMP5802400
15F (1.0 × 10−2)CsOH⋅H2O (2)NMP3101700
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cheong, Y.-J.; Lee, M.-h.; Byeon, H.; Park, J.; Yu, S.; Jang, H.-Y. Iridium(triNHC)-Catalyzed Transfer Hydrogenation of Glycerol Carbonate without Exogenous Reductants. Catalysts 2022, 12, 656. https://doi.org/10.3390/catal12060656

AMA Style

Cheong Y-J, Lee M-h, Byeon H, Park J, Yu S, Jang H-Y. Iridium(triNHC)-Catalyzed Transfer Hydrogenation of Glycerol Carbonate without Exogenous Reductants. Catalysts. 2022; 12(6):656. https://doi.org/10.3390/catal12060656

Chicago/Turabian Style

Cheong, Yeon-Joo, Mi-hyun Lee, Heemin Byeon, Jiyong Park, Sungju Yu, and Hye-Young Jang. 2022. "Iridium(triNHC)-Catalyzed Transfer Hydrogenation of Glycerol Carbonate without Exogenous Reductants" Catalysts 12, no. 6: 656. https://doi.org/10.3390/catal12060656

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