Next Article in Journal
Evolutionary Analysis of Six Gene Families Part of the Reactive Oxygen Species (ROS) Gene Network in Three Brassicaceae Species
Previous Article in Journal
Association of Alpha-Crystallin with Human Cortical and Nuclear Lens Lipid Membrane Increases with the Grade of Cortical and Nuclear Cataract
Previous Article in Special Issue
Influence of the Type of Nanofillers on the Properties of Composites Used in Dentistry and 3D Printing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multifunctional Biomass-Based Ionic Liquids/CuCl-Catalyzed CO2-Promoted Hydration of Propargylic Alcohols: A Green Synthesis of α-Hydroxy Ketones

1
State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
2
School of Material Science and Engineering, Wuhan University of Technology, Wuhan 430070, China
3
Research School of Chemical and Biomedical Technologies, National Research Tomsk Polytechnic University, Lenin Avenue 30, 634050 Tomsk, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(3), 1937; https://doi.org/10.3390/ijms25031937
Submission received: 2 January 2024 / Revised: 31 January 2024 / Accepted: 3 February 2024 / Published: 5 February 2024

Abstract

:
α-Hydroxy ketones are a class of vital organic skeletons that generally exist in a variety of natural products and high-value chemicals. However, the traditional synthetic route for their production involves toxic Hg salts and corrosive H2SO4 as catalysts, resulting in harsh conditions and the undesired side reaction of Meyer–Schuster rearrangement. In this study, CO2-promoted hydration of propargylic alcohols was achieved for the synthesis of various α-hydroxy ketones. Notably, this process was catalyzed using an environmentally friendly and cost-effective biomass-based ionic liquids/CuCl system, which effectively eliminated the side reaction. The ionic liquids utilized in this system are derived from natural biomass materials, which exhibited recyclability and catalytic activity under 1 bar of CO2 pressure without volatile organic solvents or additives. Evaluation of the green metrics revealed the superiority of this CuCl/ionic liquid system in terms of environmental sustainability. Further mechanistic investigation attributed the excellent performance to the ionic liquid component, which exhibited multifunctionality in activating substrates, CO2 and the Cu component.

Graphical Abstract

1. Introduction

Nowadays, the excessive emission of CO2 has led to a range of environmental and social problems such as global warming, glaciers melting and sea levels rising [1,2,3,4]. As a result, the effective and efficient management of CO2 has become an urgent concern for scientists and engineers. From the perspective of synthetic chemistry, CO2 is regarded as an ideal substitute for traditional phosgene and carbon monoxide [5,6,7,8] due to its wide availability, low cost, easy accessibility, non-toxicity and environmental friendliness [9]. This has prompted the exploration of CO2 for the production of fine chemicals, as it holds the potential to not only create economic value but also mitigate the greenhouse effect [10]. However, due to the thermodynamic stability and kinetic inertness of CO2, its effective activation still remains a significant challenge. Consequently, the exploration of effective catalysts for CO2 activation and the designation of feasible reaction routes for CO2 conversion have accordingly emerged as critical focal points in the pursuit of CO2 utilization.
In recent years, considerable progress has been achieved in this area [11,12,13,14,15,16,17,18,19], including effective routes for utilizing CO2 in the production of methyl urea, urea, salicylic acid, organic carbonate, methanol, polycarbonate, etc. Among them, α-hydroxy ketones are crucial organic skeletons that generally exist in a variety of natural products and are frequently used as synthetic precursors for high-value chemicals [20,21]. The hydration of propargyl alcohols was an ideal method for the production of α-hydroxy ketones due to its 100% atom economy and the accessibility of diverse starting materials. However, the direct hydration of propargyl alcohols typically required catalysts involving strong acids like H2SO4 [22,23] and rare, toxic metal salts such as Au [24,25,26,27,28,29,30,31,32,33,34], Ag [35,36,37,38,39,40] and Ru [41], which resulted in harsh conditions and the undesired side reaction of the Meyer–Schuster rearrangement [42]. Based on this, the indirect hydration of propargyl alcohols has emerged, in which the cyclization of CO2 and propargyl alcohols firstly occurred to obtain the α-alkylidene cyclic carbonates, followed by an in situ hydration of these carbonates to give the desired α-hydroxy ketones [43]. This CO2-promoted indirect hydration generally proceeded under basic conditions, thus eliminating the Meyer–Schuster rearrangement in essence. Moreover, the reaction conditions were relatively milder than those for the direct process and thus have attracted great attention from researchers and engineers.
Over the past decade, numerous catalysts have been investigated for this CO2-promoted indirect hydration. In 2014, Jiang et al. introduced a AgOAc/1,8-diazabicyclo [5.4.0]undec-7-ene (DBU) system, which successfully catalyzed this CO2-promoted hydration of internal and terminal propargyl alcohols, yielding the target products in high yield [44]. However, this system necessitated a high loading of Ag salts, traditional volatile solvents and strong bases. Moreover, its working pressure reached 20 bar of CO2. In 2015, Liu et al. presented a range of task-specific ionic liquids (ILs) for this hydration, operating effectively under 1 to 10 bar of pressure and offering a recycling and reusing ability [45]. In 2019, He et al. developed a Cu2O/DBU catalytic system composed of Cu2O (20 mol%), DBU (50 mol%) and phosphine ligands (20 mol%), which efficiently converted various types of terminal propargyl alcohols into the corresponding target products in CH3CN under 1 bar of CO2 [46]. In 2020, Yuan et al. developed a AgOAc/ILs system for this reaction, which operated under atmospheric CO2 pressure and solvent-free conditions to produce α-hydroxy ketones [47]. Furthermore, they established a Zn-based catalytic system that exhibited excellent catalytic activity for the target reaction under simulated flue gases, with the Zn species generated from pigment wastes [48]. More recently, two heterogeneous catalysts, namely a silver-anchored porous aromatic framework catalyst (Ag@PAF-DAB) and an amino-functionalized organic polymer Cu@Co-PIL-N4 loaded with highly dispersed CuI, have also been reported [49,50].
In the context of the aforementioned work, IL-involved systems have emerged as pivotal roles. ILs are composed of anions and cations with melting points below 100 °C, which have been widely applied in CO2 capture and catalytic conversion due to their advantages of designability, stability and catalytic activity [51]. However, the utilization of ILs for the CO2-promoted hydration of propargyl alcohols remains relatively uncommon and costly to date. Furthermore, some IL catalysts still required elevated CO2 pressure to reach high activity, which largely limited their further applications.
Biomass compounds are natural organic substances such as cellulose, wood chips and fructose. Nowadays, these materials have been utilized in numerous emerging areas due to their greenness, abundance, renewability, etc. [52,53]. These compounds can be further disassembled into biomass-based platform compounds [54,55], such as the 12 biomass molecules of succinic acid, 2,5-furan dicarboxylic acid, 3-hydroxypropionic acid, itaconic acid and levulinic acid, proposed by the US Department of Energy [56]. These derived compounds typically contain carboxylic or hydroxyl groups, indicating their great potential for ionization and application as the anions of ILs. Importantly, ILs containing carboxylic or hydroxyl ions have been proven to exhibit significant interaction with CO2 molecules [57]. As a result, the biomass-based ILs have the potential to demonstrate catalytic activity for CO2 capture and activation. In addition, these ILs are derived from biomass, which indicates their economical, renewable and eco-friendly nature, aligning with the requirements of modern green and sustainable development.
Herein, a series of biomass-based ILs were designed and synthesized, which were further combined with the economical CuCl for the catalysis of the CO2-promoted hydration reaction of propargyl alcohols. Particularly, this catalytic system worked under 0.1 MPa of CO2 with a low metal loading of CuCl (1 mol%) without any traditional organic solvents or ligands. Further evaluation of the green metrics revealed the superiority of this CuCl/IL system in terms of environmental sustainability.

2. Results

In this section, the catalytic performance of various biomass-based ILs for the CO2-promoted reaction was investigated, with the employment of 2-methyl-3-butyn-2-ol (1a) as the initial substrate (Table 1). The screening of ILs commenced with the blank experiments, demonstrating that the reaction could not proceed without catalysts (entry 1). Subsequently, it was observed that neither the metal salts nor the ILs alone could catalyze the target reaction (entries 2, 3). Notably, the absence of CO2 hindered the target reaction, indicating the crucial role of CO2 (entry 4). After the blank experiments, the investigation of the catalytic activity of different ILs was performed with the metal salt component fixed as CuCl. The screening was initially focused on the optimization of various anions (Figure 1), derived from levulinic acid (Lev), lactic acid (La), itaconic acid (ITa) and succinic acid (Sa). Experimental results revealed that [Lev] obtained the highest yield (entries 5–8), which was consequently identified as the optimal anion. Subsequently, the effects of cations (Figure 1) on the activity of ILs were explored, with the catalytic performance order revealed as [C2C1im] > [N4444] > [P4444] > [C4C1im] > [DBUH] > [DBNH] (entries 5, 9–13). The slight difference between [C2C1im] and [C4C1im] cations in catalytic activity can be attributed to the physical properties of the corresponding ILs. Generally, [C4C1im][Lev] exhibited higher viscosity than [C2C1im][Lev], resulting in a thicker reaction system that was more difficult to blend and stir. Consequently, the best choice of ILs was identified as [C2C1im][Lev]. With the optimal IL in hand, the metal salts in the catalytic system were further explored, which was mainly focused on economical Cu salts due to their inherent affinity to triple bonds, such as CuCl, CuBr, CuI, Cu2O, Cu2S, CuCl2, Cu(OAc)2, and CuSO4 (entries 13–20). It was found that both Cu (I) and Cu (II) salts exhibited considerable catalytic activity towards the target reaction, with CuCl achieving the highest yield of 95% (entry 13). Consequently, the optimal catalytic system was determined as CuCl/[ C2C1im][Lev].
After identifying the optimal catalytic system as CuCl/[C2C1im][Lev], the reaction conditions were subsequently explored (Table 2). The amount of [C2C1im][Lev] was gradually increased from 0.5 to 1 equiv., resulting in the highest yield of 95% (entries 1–3). A similar trend was observed for the amount of CuCl, with the yields increasing as the amount of CuCl rose from 0.25 to 1 mol%. (entries 3–6). Subsequent investigations focused on the reaction temperature. An increase in reaction temperature from 40 to 80 °C significantly improved the catalytic yield from 5% to 95%. However, at a higher temperature of 100 °C, the reaction could not proceed further, leading to the identification of 80 °C as the optimal temperature (entries 3, 7–9). Furthermore, the influence of reaction time was investigated, and the yield increased gradually as the reaction time extended (entries 3, 10–12). At a reaction time of 12 h, the yield of the product reached 95% (entry 3). Since the target reaction proceeded smoothly under 1 bar of CO2, experiments for higher pressure were not performed. Consequently, the final reaction conditions were determined as CuCl (1 mol%), [C2C1im][Lev] (1 equiv.), 80 °C, CO2 (0.1 MPa) and 12 h.
After determining the reaction conditions, the substrate scope of the CuCl/[C2C1im][Lev] system was explored (Table 3). The experimental results revealed that the majority of tertiary propargylic alcohols with various substituents effectively produced the corresponding products (1a1g). Notably, the steric effects of the substituents significantly influenced the reactivity of substrates during the formation process of α-hydroxy ketones. Substrates with less sterically hindered substituted groups such as methyl and ethyl provided high yields of 85–95% within 12 h. However, for the propargylic alcohol 1g containing a bulky phenyl group, the yield was only 33% under the same conditions, which could be improved via extending the reaction time. Additionally, attempts were made to react primary and secondary propargylic alcohols (1h, 1i), but the corresponding products could not be produced. This might be attributed to the lack of gem-dialkyl effects in these substrates, resulting in the failure of cyclizing CO2 and propargylic alcohol [58,59]. Subsequently, the recyclability of the CuCl/[C2C1im][Lev] system was investigated. After being recycled and reused three times, the catalytic system could still catalyze the target reaction to produce the α-hydroxy ketones with a yield of 90%, demonstrating its considerable stability and recyclability.
In addition to substrate scope and recyclability, the greenness of the reaction process is another important aspect of modern sustainable development, which can be quantified by utilizing the “green metrics”, including atom economy (AE), E-factor, carbon efficiency (CE), reaction mass efficiency (RME), mass intensity (MI) and mass productivity (MP) [60]. These well-defined and objective metrics provide quantitative standards for the greenness evaluation (Part 2, supporting information). In this context, the green metrics of the CuCl/[C2C1im][Lev] system and the calculable systems reported by other researchers were compared, based on the hydration of 1a (Table 4). Upon evaluating a total of six green metrics, the CuCl/[C2C1im][Lev] system provided superior values in four aspects (AE, E-factor, MI, MP). This result indicated that the CuCl/[C2C1im][Lev]-catalyzed hydration of propargyl alcohols was a relatively greener process.

3. Discussion

3.1. Identification of Tandem Mechanism

Based on previous reports [47], the CO2-promoted hydration of propargyl alcohols may proceed via a two-step tandem reaction mechanism. Substrates and CO2 may first undergo cyclization to form α-alkylidene cyclic carbonates, followed by the in situ hydration of these carbonates with the release of CO2 during the process. To investigate whether the CuCl/[C2C1im][Lev]-catalyzed reaction aligns with this proposed mechanism, a series of control experiments were conducted. Initially, 1a, CO2 and CuCl/[C2C1im][Lev] were introduced into the system under the optimal reaction conditions, with the omission of H2O. After 12 h, the reaction was terminated, and the sample was analyzed using 1H NMR. Comparing the spectrum of the reaction mixture with that of pure α-alkylidene cyclic carbonate, the characteristic peaks of α-alkylidene cyclic carbonates appeared on the spectrum of the reaction mixture (Figure 2a), indicating the generation of these carbonates in this process. Subsequently, the pure α-alkylidene cyclic carbonates were added with H2O and allowed to react for 12 h under the catalysis of [C2C1im][Lev]. Upon completion of the reaction, the sample was analyzed using 1H NMR. The result showed that the desired α-hydroxy ketones were successfully obtained (Figure 2b). These experiments demonstrated that the CuCl/[C2C1im][Lev]-catalyzed CO2-promoted hydration of propargyl alcohols followed the proposed tandem reaction mechanism, and the ILs act as the pivotal catalyst for the hydration process.

3.2. Activation of the Propargylic Alcohols

Upon identifying the tandem mechanism of the target reaction, the first cyclization step was further studied. In this step, the activation of the hydroxyl group in propargyl alcohols was quite crucial, which would initiate the whole catalytic process. Therefore, the investigation focused on identifying the component responsible for this crucial activation. Typically, this activation could be indicated via the shape and chemical shift of the hydroxyl signal peak in the 1H NMR spectrum. For pure 1a, the hydroxyl proton exhibited a distinct characteristic peak at δ = 5.29 ppm (Figure 3a), indicating its inactivated state. However, upon the addition of DBU, a well-established organic base known for its effective hydroxyl group activation, the signal transformed into a broad peak with a different chemical shift (Figure 3b). This represented the activated state of the hydroxyl proton. Subsequently, the two components of the catalytic system were successively scrutinized. The addition of CuCl to 1a did not produce any discernible change in the 1H NMR results (Figure 3c), suggesting that CuCl was not capable of activating 1a. Conversely, the combination of [C2C1im][Lev] and 1a led to a broad and shifted hydroxyl peak similar to that in the DBU/1a system (Figure 3d). This result implied that [C2C1im][Lev] played a pivotal role in the activation of the hydroxyl group in 1a.

3.3. Generation of NHC-CO2 Adducts

It has been reported that under basic conditions, imidazole compounds might interact with CO2, resulting in the generation of free N-heterocyclic carbenes (NHCs) that are capable of CO2 capture and activation via the formation of NHC-CO2 adducts (Scheme 1) [62,63]. In the aforementioned investigations, [C2C1im][Lev] was identified to provide suitable basic conditions for the activation of the hydroxyl group. Therefore, subsequent exploration focused on whether this imidazole IL could produce NHC-CO2 adducts. In this study, CO2 was introduced into [C2C1im][Lev], and the mixture was allowed to stir for 12 h. It could be observed that the solution became turbid gradually. Upon completion, the solution was analyzed using 13C NMR. In the spectrum, a new signal peak at δ = 154.60 ppm was observed (Figure 4a), which was consistent with the characteristic peak of CO2 adducts reported in the literature [63], indicating the successful formation of NHC-CO2 adducts.

3.4. Interaction between [C2C1im][Lev] and Cu Species

Apart from the function in activating hydroxyl groups and generating NHC-CO2 adducts, the [C2C1im][Lev] component was further identified to improve the catalytic activity of CuCl via the interaction between [Cu] and carbonyl groups in the [Lev] anions. In this context, an analogous IL to [C2C1im][Lev] was synthesized, differing only in the absence of a carbonyl group in its anion structure (Scheme 2, entry 1). Subsequent experiments were designed to compare the catalytic performance of these two ILs. The results illustrated a significant decrease in catalytic yields in the absence of carbonyl groups, demonstrating the indispensable roles of carbonyls in ILs. Further evidence of the interaction between the carbonyl groups and [Cu] was revealed via the 1H NMR of the CuCl/[C2C1im][Lev] mixture (Figure 5). At 80 °C, significant interactions were observed as the signals of CH2 groups adjacent to the carboxyl groups (dashed part, Figure 5) shifted and broadened. This contrasts with the sharp triplet peaks observed at 25 °C, suggesting that the interaction at lower temperature was less significant. This observation provides a plausible explanation that the CuCl/[C2C1im][Lev] system exhibited better activity at a relatively higher temperature.

3.5. Proposed Mechanism for the Catalytic Process

Based on the aforementioned experiments and discussions [44,57,64,65,66], the following mechanism for the CO2-promoted hydration of propargylic alcohols catalyzed via the CuCl/[C2C1im][Lev] system was proposed, which could be outlined in two steps (Scheme 3). Firstly, [Lev] activates the hydroxy group of 1a, thereby enhancing the nucleophilicity of the hydroxy oxygen and facilitating its subsequent attack on the CO2 molecule. Simultaneously, the Cu species coordinates with the unsaturated triple bond of 1a, leading to the formation of the transition state (TS). In the next stage, the negative hydroxyl oxygen atom bonds with the positive carbon center of CO2, incorporating the inert CO2 molecule into the organic skeletons, resulting in the formation of intermediate I. Subsequently, the negative oxygen of the CO2 moiety continues to attack the triple bond, activated via the Cu species, leading to intramolecular cyclization and the formation of intermediate II. Finally, the proton is transferred back to the organic skeletons, resulting in the generation of the key α-alkylidene carbonate (intermediate III). In the second step, H2O initiates a nucleophilic attack on intermediate III with the catalysis of the basic [Lev]. This leads to the ring-opening reaction and the generation of intermediate IV. Afterwards, this intermediate undergoes keto–enol isomerization, subsequently releasing the CO2 molecule and yielding the final target product 2a.

4. Materials and Methods

The series of biomass-based ILs used in the experiments were synthesized according to the reported literature (Part 1, supporting information) [57,67,68]. Unless otherwise specified, all the propargyl alcohol substrates (98%) and biomass acids (95%) used in the experiments were purchased from Aladdin (Shanghai, China), TCI (Tokyo, Japan), Sigma-Aldrich (Shanghai, China), Macklin (Shanghai, China), Alfa (Shanghai, China), etc. and directly used without further purification and drying. The purity of the CO2 used for purging and reacting was 99.9%, supplied by Wuhan Xiangyun Industry and Trade Co., Ltd., (Wuhan, China).
The 1H NMR spectra were recorded on a Bruker Avance III HD 500 MHz spectrometer, with the internal standard TMS (δ = 0 ppm) serving as the reference. Meanwhile, the 13C NMR spectra were recorded at 126 MHz in CDCl3 (δ = 77.23 ppm) or DMSO-d6 (δ = 39.50 ppm), with the solvent peaks as the internal references. The data were given as chemical shifts (ppm) and coupling constants (Hz), respectively.

4.1. The CO2-Promoted Hydration of Propargylic Alcohols

CuCl (0.025 mmol), 1-ethyl-3-methylimidazolium levulinic ([C2C1im][Lev], 2.5 mmol), propargylic alcohols (2.5 mmol) and H2O (5 mmol) were added into a Schlenk tube. Subsequently, the system was purged three times with CO2 and then stirred at 80 °C under 0.1 MPa of CO2 for the required time. When the reaction was completed, the mixture was extracted with diethyl ether (3 × 15 mL). The upper organic phases were concentrated under a vacuum to give the crude products, which were further purified via column chromatography on silica gel using petroleum ether/ethyl acetate (v/v, 100:1–20:1) as the eluent.

4.2. Procedures for Recycling the Catalytic System

After the reaction was completed, the mixture was extracted three times with diethyl ether (3 × 15 mL). The lower layer was then dried under a vacuum for 4 h to totally remove the residual solvents, reactants and products. After drying, the catalytic system could be reused for the next round.

5. Conclusions

In summary, a CuCl/[C2C1im][Lev] system was developed for the CO2-promoted hydration of propargylic alcohols under 1 bar of CO2 pressure without the use of organic volatile solvents or additives. The system demonstrated high activity, a wide substrate scope and significant recyclability. Notably, the catalytic process was identified as superior in terms of green metric evaluation. A comprehensive mechanistic investigation revealed that the exceptional performance of this system can be attributed to the triple function of the [C2C1im][Lev] component, which encompasses the activation of substrates, generation of NHC-CO2 adducts and interaction with the Cu component.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25031937/s1.

Author Contributions

Conceptualization, Y.Y. and F.V.; methodology, S.Z.; software, K.D.; validation, Y.Y., S.Z. and K.D.; formal analysis, Y.X.; investigation, K.G.; resources, C.C.; data curation, S.C.; writing—original draft preparation, Y.Y.; writing—review and editing, F.V.; visualization, S.Z.; supervision, D.C. and F.V.; project administration, F.V.; funding acquisition, Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

The project was supported by the National Natural Science Foundation of China (no. 22102127).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Oerlemans, J. Glaciers as indicators of a carbon dioxide warming. Nature 1986, 320, 607–609. [Google Scholar] [CrossRef]
  2. McNutt, M. Time’s up, CO2. Science 2019, 365, 411. [Google Scholar] [CrossRef] [PubMed]
  3. Mac Dowell, N.; Reiner, D.M.; Haszeldine, R.S. Comparing approaches for carbon dioxide removal. Joule 2022, 6, 2233–2239. [Google Scholar] [CrossRef]
  4. Frölicher, T.L.; Fischer, E.M.; Gruber, N. Marine heatwaves under global warming. Nature 2018, 560, 360–364. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, Z.-Z.; He, L.-N.; Gao, J.; Liu, A.-H.; Yu, B. Carbon dioxide utilization with C–N bond formation: Carbon dioxide capture and subsequent conversion. Energy Environ. Sci. 2012, 5, 6602–6639. [Google Scholar] [CrossRef]
  6. Huang, K.; Sun, C.-L.; Shi, Z.-J. Transition-metal-catalyzed C–C bond formation through the fixation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 2435–2452. [Google Scholar] [CrossRef] [PubMed]
  7. He, M.; Sun, Y.; Han, B. Green Carbon Science: Scientific Basis for Integrating Carbon Resource Processing, Utilization, and Recycling. Angew. Chem. Int. Ed. 2013, 52, 9620–9633. [Google Scholar] [CrossRef]
  8. Bao, J.; Yang, G.; Yoneyama, Y.; Tsubaki, N. Significant Advances in C1 Catalysis: Highly Efficient Catalysts and Catalytic Reactions. ACS Catal. 2019, 9, 3026–3053. [Google Scholar] [CrossRef]
  9. Kleij, A.W.; North, M.; Urakawa, A. CO2 Catalysis. ChemSusChem 2017, 10, 1036–1038. [Google Scholar] [CrossRef]
  10. Mac Dowell, N.; Fennell, P.S.; Shah, N.; Maitland, G.C. The role of CO2 capture and utilization in mitigating climate change. Nat. Clim. Change 2017, 7, 243–249. [Google Scholar] [CrossRef]
  11. Hou, S.-L.; Dong, J.; Zhao, B. Formation of C-X Bonds in CO2 Chemical Fixation Catalyzed by Metal−Organic Frameworks. Adv. Mater. 2020, 32, 1806163. [Google Scholar] [CrossRef]
  12. 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]
  13. Riemer, D.; Schilling, W.; Goetz, A.; Zhang, Y.; Gehrke, S.; Tkach, I.; Hollóczki, O.; Das, S. CO2-Catalyzed Efficient Dehydrogenation of Amines with Detailed Mechanistic and Kinetic Studies. ACS Catal. 2018, 8, 11679–11687. [Google Scholar] [CrossRef]
  14. Riemer, D.; Mandaviya, B.; Schilling, W.; Götz, A.C.; Kühl, T.; Finger, M.; Das, S. CO2-Catalyzed Oxidation of Benzylic and Allylic Alcohols with DMSO. ACS Catal. 2018, 8, 3030–3034. [Google Scholar] [CrossRef]
  15. Hirapara, P.; Riemer, D.; Hazra, N.; Gajera, J.; Finger, M.; Das, S. CO2-assisted synthesis of non-symmetric α-diketones directly from aldehydes via C–C bond formation. Green Chem. 2017, 19, 5356–5360. [Google Scholar] [CrossRef]
  16. Riemer, D.; Hirapara, P.; Das, S. Chemoselective Synthesis of Carbamates using CO2 as Carbon Source. ChemSusChem 2016, 9, 1916–1920. [Google Scholar] [CrossRef] [PubMed]
  17. Aresta, M.; Dibenedetto, A.; Angelini, A. Catalysis for the Valorization of Exhaust Carbon: From CO2 to Chemicals, Materials, and Fuels. Technological Use of CO2. Chem. Rev. 2014, 114, 1709–1742. [Google Scholar] [CrossRef] [PubMed]
  18. Appel, A.M.; Bercaw, J.E.; Bocarsly, A.B.; Dobbek, H.; DuBois, D.L.; Dupuis, M.; Ferry, J.G.; Fujita, E.; Hille, R.; Kenis, P.J.A.; et al. Frontiers, Opportunities, and Challenges in Biochemical and Chemical Catalysis of CO2 Fixation. Chem. Rev. 2013, 113, 6621–6658. [Google Scholar] [CrossRef] [PubMed]
  19. Sakakura, T.; Choi, J.-C.; Yasuda, H. Transformation of Carbon Dioxide. Chem. Rev. 2007, 107, 2365–2387. [Google Scholar] [CrossRef]
  20. Otto, A.; Grube, T.; Schiebahn, S.; Stolten, D. Closing the loop: Captured CO2 as a feedstock in the chemical industry. Energy Environ. Sci. 2015, 8, 3283–3297. [Google Scholar] [CrossRef]
  21. Yang, K.; Zhang, F.; Fang, T.; Li, C.; Li, W.; Song, Q. Passerini-type reaction of boronic acids enables α-hydroxyketones synthesis. Nat. Commun. 2021, 12, 441. [Google Scholar] [CrossRef]
  22. Kutscheroff, M. Ueber die Einwirkung der Kohlenwasserstoffe der Acetylenreihe auf Quecksilberoxyd und dessen Salze. Ber. Dtsch. Chem. Ges. 1884, 17, 13–29. [Google Scholar] [CrossRef]
  23. Kutscheroff, M. Ueber eine neue Methode direkter Addition von Wasser (Hydratation) an die Kohlenwasserstoffe der Acetylenreihe. Ber. Dtsch. Chem. Ges. 1881, 14, 1540–1542. [Google Scholar] [CrossRef]
  24. Xing, Y.; Zhang, M.; Ciccarelli, S.; Lee, J.; Catano, B. AuIII-Catalyzed Formation of α-Halomethyl Ketones from Terminal Alkynes. Eur. J. Org. Chem. 2017, 2017, 781–785. [Google Scholar] [CrossRef]
  25. Weerasiri, K.C.; Chen, D.; Wozniak, D.I.; Dobereiner, G.E. Internal Alkyne Regio- and Chemoselectivity using a Zwitterionic N-Heterocyclic Carbene Gold Catalyst in a Silver-Free Alkyne Hydration Reaction. Adv. Synth. Catal. 2016, 358, 4106–4113. [Google Scholar] [CrossRef]
  26. Gatto, M.; Belanzoni, P.; Belpassi, L.; Biasiolo, L.; Del Zotto, A.; Tarantelli, F.; Zuccaccia, D. Solvent-, Silver-, and Acid-Free NHC-Au-X Catalyzed Hydration of Alkynes. The Pivotal Role of the Counterion. ACS Catal. 2016, 6, 7363–7376. [Google Scholar] [CrossRef]
  27. Ebule, R.E.; Malhotra, D.; Hammond, G.B.; Xu, B. Ligand Effects in the Gold Catalyzed Hydration of Alkynes. Adv. Synth. Catal. 2016, 358, 1478–1481. [Google Scholar] [CrossRef]
  28. Xu, Y.; Hu, X.; Shao, J.; Yang, G.; Wu, Y.; Zhang, Z. Hydration of alkynes at room temperature catalyzed by gold(i) isocyanide compounds. Green Chem. 2015, 17, 532–537. [Google Scholar] [CrossRef]
  29. Liang, S.; Jasinski, J.; Hammond, G.B.; Xu, B. Supported Gold Nanoparticle-Catalyzed Hydration of Alkynes under Basic Conditions. Org. Lett. 2015, 17, 162–165. [Google Scholar] [CrossRef]
  30. Ibrahim, H.; de Frémont, P.; Braunstein, P.; Théry, V.; Nauton, L.; Cisnetti, F.; Gautier, A. Water-Soluble Gold–N-Heterocyclic Carbene Complexes for the Catalytic Homogeneous Acid- and Silver-Free Hydration of Hydrophilic Alkynes. Adv. Synth. Catal. 2015, 357, 3893–3900. [Google Scholar] [CrossRef]
  31. Chen, T.; Cai, C. Catalytic hydration of alkynes to ketones by a salen–gold(III) complex. Catal. Commun. 2015, 65, 102–104. [Google Scholar] [CrossRef]
  32. Xie, L.; Wu, Y.; Yi, W.; Zhu, L.; Xiang, J.; He, W. Gold-Catalyzed Hydration of Haloalkynes to α-Halomethyl Ketones. J. Org. Chem. 2013, 78, 9190–9195. [Google Scholar] [CrossRef] [PubMed]
  33. Oonishi, Y.; Gómez-Suárez, A.; Martin, A.R.; Nolan, S.P. Hydrophenoxylation of Alkynes by Cooperative Gold Catalysis. Angew. Chem. Int. Ed. 2013, 52, 9767–9771. [Google Scholar] [CrossRef] [PubMed]
  34. Leyva, A.; Corma, A. Isolable Gold(I) Complexes Having One Low-Coordinating Ligand as Catalysts for the Selective Hydration of Substituted Alkynes at Room Temperature without Acidic Promoters. J. Org. Chem. 2009, 74, 2067–2074. [Google Scholar] [CrossRef] [PubMed]
  35. Santhi, J.; Baire, B. Carbonyl Directed Regioselective Hydration of Alkynes under Ag-Catalysis. ChemistrySelect 2017, 2, 4338–4342. [Google Scholar] [CrossRef]
  36. Dong, Q.; Li, N.; Qiu, R.; Wang, J.; Guo, C.; Xu, X. Silver-containing microemulsion as a high-efficient and recyclable catalytic system for hydration of alkynes. J. Organomet. Chem. 2015, 799–800, 122–127. [Google Scholar] [CrossRef]
  37. Chen, Z.-W.; Ye, D.-N.; Ye, M.; Zhou, Z.-G.; Li, S.-H.; Liu, L.-X. AgF/TFA-promoted highly efficient synthesis of α-haloketones from haloalkynes. Tetrahedron Lett. 2014, 55, 1373–1375. [Google Scholar] [CrossRef]
  38. Chen, Z.-W.; Ye, D.-N.; Qian, Y.-P.; Ye, M.; Liu, L.-X. Highly efficient AgBF4-catalyzed synthesis of methyl ketones from terminal alkynes. Tetrahedron 2013, 69, 6116–6120. [Google Scholar] [CrossRef]
  39. Venkateswara Rao, K.T.; Sai Prasad, P.S.; Lingaiah, N. Solvent-free hydration of alkynes over a heterogeneous silver exchanged silicotungstic acid catalyst. Green Chem. 2012, 14, 1507–1514. [Google Scholar] [CrossRef]
  40. Das, R.; Chakraborty, D. AgOTf catalyzed hydration of terminal alkynes. Appl. Organomet. Chem. 2012, 26, 722–726. [Google Scholar] [CrossRef]
  41. Grotjahn, D.B.; Lev, D.A. A General Bifunctional Catalyst for the Anti-Markovnikov Hydration of Terminal Alkynes to Aldehydes Gives Enzyme-Like Rate and Selectivity Enhancements. J. Am. Chem. Soc. 2004, 126, 12232–12233. [Google Scholar] [CrossRef]
  42. Banerjee, S.; Ambegave, S.B.; Mule, R.D.; Senthilkumar, B.; Patil, N.T. Gold-Catalyzed Alkynylative Meyer–Schuster Rearrangement. Org. Lett. 2020, 22, 4792–4796. [Google Scholar] [CrossRef]
  43. Yoshida, S.; Fukui, K.; Kikuchi, S.; Yamada, T. Silver-Catalyzed Enantioselective Carbon Dioxide Incorporation into Bispropargylic Alcohols. J. Am. Chem. Soc. 2010, 132, 4072–4073. [Google Scholar] [CrossRef] [PubMed]
  44. He, H.; Qi, C.; Hu, X.; Guan, Y.; Jiang, H. Efficient synthesis of tertiary α-hydroxy ketones through CO2-promoted regioselective hydration of propargylic alcohols. Green Chem. 2014, 16, 3729–3733. [Google Scholar] [CrossRef]
  45. Zhao, Y.; Yang, Z.; Yu, B.; Zhang, H.; Xu, H.; Hao, L.; Han, B.; Liu, Z. Task-specific ionic liquid and CO2-cocatalysed efficient hydration of propargylic alcohols to α-hydroxy ketones. Chem. Sci. 2015, 6, 2297–2301. [Google Scholar] [CrossRef] [PubMed]
  46. Zhou, Z.-H.; Zhang, X.; Huang, Y.-F.; Chen, K.-H.; He, L.-N. Synthesis of α-hydroxy ketones by copper(I)-catalyzed hydration of propargylic alcohols: CO2 as a cocatalyst under atmospheric pressure. Chin. J. Catal. 2019, 40, 1345–1351. [Google Scholar] [CrossRef]
  47. Li, D.; Gong, Y.; Du, M.; Bu, C.; Chen, C.; Chaemcheun, S.; Hu, J.; Zhang, Y.; Yuan, Y.; Verpoort, F. CO2-Promoted Hydration of Propargylic Alcohols: Green Synthesis of α-Hydroxy Ketones by an Efficient and Recyclable AgOAc/Ionic Liquid System. ACS Sustain. Chem. Eng. 2020, 8, 8148–8155. [Google Scholar] [CrossRef]
  48. Zhang, Y.; Hu, J.; Xu, Y.; Yan, X.; Zhang, S.; Duan, K.; Chen, C.; Yuan, Y.; Verpoort, F. CO2-induced dissolution of ZnO into ionic liquids and its catalytic application for the hydration of propargylic alcohols. Appl. Catal. B 2022, 310, 121270. [Google Scholar] [CrossRef]
  49. Guo, X.-X.; Cai, Z.-T.; Muhammad, Y.; Zhang, F.-L.; Wei, R.-P.; Gao, L.-J.; Xiao, G.-M. Silver-anchored porous aromatic framework for efficient conversion of propargylic alcohols with CO2 at ambient pressure. Chin. Chem. Lett. 2023, 34, 107740. [Google Scholar] [CrossRef]
  50. Guo, X.; Zhang, F.; Muhammad, Y.; Yang, Z.; Wei, R.; Gao, L.; Xiao, G. Amino-functionalized organic polymer loaded with highly dispersed CuI for efficient catalytic conversion of CO2 with PA. Microporous Mesoporous Mater. 2023, 352, 112507. [Google Scholar] [CrossRef]
  51. Simon, N.M.; Zanatta, M.; dos Santos, F.P.; Corvo, M.C.; Cabrita, E.J.; Dupont, J. Carbon Dioxide Capture by Aqueous Ionic Liquid Solutions. ChemSusChem 2017, 10, 4927–4933. [Google Scholar] [CrossRef] [PubMed]
  52. Zhao, H.-Y.; Dong, W.-X.; Deng, Y.; Chen, L.-F.; Zhao, C.-F.; Zhang, C.-L.; Zhou, J.; Qu, Y.-F.; Li, Y.-S.; Li, D.-J.; et al. Biomass-based biomimetic-oriented Janus nanoarchitecture for efficient heavy-metal enrichment and interfacial solar water sanitation. Interdiscip. Mater. 2022, 1, 537–547. [Google Scholar] [CrossRef]
  53. Li, T.; Chen, C.; Brozena, A.H.; Zhu, J.Y.; Xu, L.; Driemeier, C.; Dai, J.; Rojas, O.J.; Isogai, A.; Wågberg, L.; et al. Developing fibrillated cellulose as a sustainable technological material. Nature 2021, 590, 47–56. [Google Scholar] [CrossRef] [PubMed]
  54. Tiong, Y.W.; Yap, C.L.; Gan, S.; Yap, W.S.P. Conversion of Biomass and Its Derivatives to Levulinic Acid and Levulinate Esters via Ionic Liquids. Ind. Eng. Chem. Res. 2018, 57, 4749–4766. [Google Scholar] [CrossRef]
  55. Bernardo, J.R.; Oliveira, M.C.; Fernandes, A.C. HReO4 as highly efficient and selective catalyst for the conversion of carbohydrates into value added chemicals. Mol. Catal. 2019, 465, 87–94. [Google Scholar] [CrossRef]
  56. Werpy, T.A.; Holladay, J.E.; White, J.F. Top Value Added Chemicals From Biomass: I. Results of Screening for Potential Candidates from Sugars and Synthesis Gas; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2004.
  57. Hu, Y.; Song, J.; Xie, C.; Wu, H.; Jiang, T.; Yang, G.; Han, B. Transformation of CO2 into α-Alkylidene Csights into the levulinate-based ionic liquid class: By CuI and Ionic Liquid with Biomass-Derived Levulinate Anion. ACS Sustain. Chem. Eng. 2019, 7, 5614–5619. [Google Scholar] [CrossRef]
  58. Grignard, B.; Ngassamtounzoua, C.; Gennen, S.; Gilbert, B.; Méreau, R.; Jerome, C.; Tassaing, T.; Detrembleur, C. Boosting the Catalytic Performance of Organic Salts for the Fast and Selective Synthesis of α-Alkylidene Cyclic Carbonates from Carbon Dioxide and Propargylic Alcohols. ChemCatChem 2018, 10, 2584–2592. [Google Scholar] [CrossRef]
  59. Kayaki, Y.; Yamamoto, M.; Ikariya, T. Stereoselective Formation of α-Alkylidene Cyclic Carbonates via Carboxylative Cyclization of Propargyl Alcohols in Supercritical Carbon Dioxide. J. Org. Chem. 2007, 72, 647–649. [Google Scholar] [CrossRef]
  60. Constable, D.J.C.; Curzons, A.D.; Cunningham, V.L. Metrics to ‘green’ chemistry—Which are the best? Green Chem. 2002, 4, 521–527. [Google Scholar] [CrossRef]
  61. Tang, M.; Zhang, F.; Zhao, Y.; Wang, Y.; Ke, Z.; Li, R.; Zeng, W.; Han, B.; Liu, Z. A CO2-mediated base catalysis approach for the hydration of triple bonds in ionic liquids. Green Chem. 2021, 23, 9870–9875. [Google Scholar] [CrossRef]
  62. Kayaki, Y.; Yamamoto, M.; Ikariya, T. N-Heterocyclic Carbenes as Efficient Organocatalysts for CO2 Fixation Reactions. Angew. Chem. Int. Ed. 2009, 48, 4194–4197. [Google Scholar] [CrossRef]
  63. Gurau, G.; Rodríguez, H.; Kelley, S.P.; Janiczek, P.; Kalb, R.S.; Rogers, R.D. Demonstration of Chemisorption of Carbon Dioxide in 1,3-Dialkylimidazolium Acetate Ionic Liquids. Angew. Chem. Int. Ed. 2011, 50, 12024–12026. [Google Scholar] [CrossRef]
  64. Zhao, Y.; Tian, L.; Qiu, J.; Li, Z.; Wang, H.; Cui, G.; Zhang, S.; Wang, J. Remarkable synergistic effect between copper(I) and ionic liquids for promoting chemical fixation of CO2. J. CO2 Util. 2017, 22, 374–381. [Google Scholar] [CrossRef]
  65. Shi, G.; Zhai, R.; Li, H.; Wang, C. Highly efficient synthesis of alkylidene cyclic carbonates from low concentration CO2 using hydroxyl and azolate dual functionalized ionic liquids. Green Chem. 2021, 23, 592–596. [Google Scholar] [CrossRef]
  66. Ouyang, L.; Tang, X.; He, H.; Qi, C.; Xiong, W.; Ren, Y.; Jiang, H. Copper-Promoted Coupling of Carbon Dioxide and Propargylic Alcohols: Expansion of Substrate Scope and Trapping of Vinyl Copper Intermediate. Adv. Synth. Catal. 2015, 357, 2556–2565. [Google Scholar] [CrossRef]
  67. Mezzetta, A.; Becherini, S.; Pretti, C.; Monni, G.; Casu, V.; Chiappe, C.; Guazzelli, L. Insights into the levulinate-based ionic liquid class: Synthesis, cellulose dissolution evaluation and ecotoxicity assessment. New J. Chem. 2019, 43, 13010–13019. [Google Scholar] [CrossRef]
  68. Deng, L.; Yue, W.; Zhang, L.; Guo, Y.; Xie, H.; Zheng, Q.; Zou, G.; Chen, P. Biobased Protic Ionic Liquids as Sustainable Solvents for Wool Keratin/Cellulose Simultaneous Dissolution: Solution Properties and Composited Membrane Preparation. ACS Sustain. Chem. Eng. 2022, 10, 2158–2168. [Google Scholar] [CrossRef]
Figure 1. The cations and anions of the ILs.
Figure 1. The cations and anions of the ILs.
Ijms 25 01937 g001
Figure 2. Comparison of 1H NMR spectra: (a) pure a-alkylidene carbonates vs. mixture of step 1; (b) pure a-hydroxy ketones vs. mixture of step 2.
Figure 2. Comparison of 1H NMR spectra: (a) pure a-alkylidene carbonates vs. mixture of step 1; (b) pure a-hydroxy ketones vs. mixture of step 2.
Ijms 25 01937 g002
Figure 3. 1H NMR of (a) pure 1a, (b) 1a/DBU (1:1), (c) 1a/CuCl (1:1), (d) 1a/[C2C1im][Lev] (1:1) in DMSO-d6.
Figure 3. 1H NMR of (a) pure 1a, (b) 1a/DBU (1:1), (c) 1a/CuCl (1:1), (d) 1a/[C2C1im][Lev] (1:1) in DMSO-d6.
Ijms 25 01937 g003
Scheme 1. Formation of NHC-CO2 adducts in the imidazole-based system.
Scheme 1. Formation of NHC-CO2 adducts in the imidazole-based system.
Ijms 25 01937 sch001
Figure 4. The detection of NHC-CO2 complex ((a): [C2C1im][Lev] + CO2; (b): pure [C2C1im][Lev]).
Figure 4. The detection of NHC-CO2 complex ((a): [C2C1im][Lev] + CO2; (b): pure [C2C1im][Lev]).
Ijms 25 01937 g004
Scheme 2. Parallel experiments employed IL1 and IL2. (Conditions: substrate: 2.5 mmol, CuCl: 1 mol%, H2O: 5 mmol, IL: 2.5 mmol, CO2: 0.1 MPa, 80 °C).
Scheme 2. Parallel experiments employed IL1 and IL2. (Conditions: substrate: 2.5 mmol, CuCl: 1 mol%, H2O: 5 mmol, IL: 2.5 mmol, CO2: 0.1 MPa, 80 °C).
Ijms 25 01937 sch002
Figure 5. 1H NMR for the mixture of CuCl and [C2C1im][Lev] at 25 °C (a), 45 °C (b), 80 °C (c).
Figure 5. 1H NMR for the mixture of CuCl and [C2C1im][Lev] at 25 °C (a), 45 °C (b), 80 °C (c).
Ijms 25 01937 g005
Scheme 3. Proposed catalytic mechanism of the CuCl/[C2C1im][Lev] system.
Scheme 3. Proposed catalytic mechanism of the CuCl/[C2C1im][Lev] system.
Ijms 25 01937 sch003
Table 1. Screening of catalytic systems a.
Table 1. Screening of catalytic systems a.
Ijms 25 01937 i001
EntryMetal SaltILYield (%) b
1- c- d0
2CuCl- d0
3- c[C2C1im][Lev]0
4 eCuCl[C2C1im][Lev]0
5CuCl[N4444][Lev]94
6CuCl[N4444]2[ITa]80
7CuCl[N4444]2[Sa]88
8CuCl[N4444][La]74
9CuCl[P4444][Lev]87
10CuCl[DBUH][Lev]75
11CuCl[DBNH][Lev]47
12CuCl[C4C1im][Lev]85
13CuCl[C2C1im][Lev]95
14CuBr[C2C1im][Lev]89
15CuI[C2C1im][Lev]66
16Cu2O[C2C1im][Lev]86
17Cu2S[C2C1im][Lev]72
18CuCl2[C2C1im][Lev]67
19Cu(OAc)2[C2C1im][Lev]80
20CuSO4[C2C1im][Lev]73
a Conditions: metal salt (1 mol%), 1a (2.5 mmol), H2O (5 mmol), ILs (2.5 mmol), 80 °C, CO2 (0.1 MPa), 12 h; b yields were determined via 1H NMR spectroscopy using 1,3,5-trimethoxybenzene as the internal standard; c without metal salt; d without IL; e absence of CO2.
Table 2. Screening of the reaction conditions a.
Table 2. Screening of the reaction conditions a.
Ijms 25 01937 i002
EntryCuCl (mol %) cIL (equiv.) cTemperature (°C)Time (h)Yield (%) b
110.5801241
210.75801262
311801295
40.751801276
50.51801251
60.251801242
71140125
811601278
9111001293
101180340
111180660
121180980
a Reaction conditions: 1a (2.5 mmol), H2O (5 mmol), CO2 (0.1 MPa). b Yields were determined via 1H NMR spectroscopy using 1,3,5-trimethoxybenzene as the internal standard. c Based on the amount of 1a.
Table 3. Screening of the substrates a.
Table 3. Screening of the substrates a.
Ijms 25 01937 i003
SubstrateProductTime (h)Yield (%) b
1aIjms 25 01937 i0042aIjms 25 01937 i0051295 (90 c)
1bIjms 25 01937 i0062bIjms 25 01937 i00712 89
1cIjms 25 01937 i0082cIjms 25 01937 i0091287
1dIjms 25 01937 i0102dIjms 25 01937 i0111285
1eIjms 25 01937 i0122eIjms 25 01937 i0131281
1fIjms 25 01937 i0142fIjms 25 01937 i01512
24
69
79
1gIjms 25 01937 i0162gIjms 25 01937 i01712
36
33
73
1hIjms 25 01937 i0182hIjms 25 01937 i019120
1iIjms 25 01937 i0202iIjms 25 01937 i021120
a Conditions: 1 (2.5 mmol), CuCl (1 mol%), H2O (5 mmol), [C2C1im][Lev] (2.5 mmol), CO2 (0.1 MPa), 80 °C; b yields were determined via 1H NMR spectroscopy using 1,3,5-trimethoxybenzene as the internal standard. c After 3 times of recycling and reuse.
Table 4. Comparison of the green metrics.
Table 4. Comparison of the green metrics.
Ref.CatalystSolventAE
(%)
E-Factor (kg/kg)CE
(%)
RME
(%)
MI
(kg/kg)
MP
(%)
2015 [45][Bu4P][Im]/10011.29278.212.28.2
2019 [46]Cu2O/DBU cyclohexyldiphenylphosphineCH3CN10010.09797.111.09.1
2021 [61][P4444][2-OP]/1004.98383.05.916.9
2023 [49]Silver-anchored porous aromatic framework
Ag@PAF-DAB
CH3CN10017.79986.318.75.4
this workCuCl/[C2C1im][Lev]/1003.09580.84.024.7
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yuan, Y.; Zhang, S.; Duan, K.; Xu, Y.; Guo, K.; Chen, C.; Chaemchuen, S.; Cao, D.; Verpoort, F. Multifunctional Biomass-Based Ionic Liquids/CuCl-Catalyzed CO2-Promoted Hydration of Propargylic Alcohols: A Green Synthesis of α-Hydroxy Ketones. Int. J. Mol. Sci. 2024, 25, 1937. https://doi.org/10.3390/ijms25031937

AMA Style

Yuan Y, Zhang S, Duan K, Xu Y, Guo K, Chen C, Chaemchuen S, Cao D, Verpoort F. Multifunctional Biomass-Based Ionic Liquids/CuCl-Catalyzed CO2-Promoted Hydration of Propargylic Alcohols: A Green Synthesis of α-Hydroxy Ketones. International Journal of Molecular Sciences. 2024; 25(3):1937. https://doi.org/10.3390/ijms25031937

Chicago/Turabian Style

Yuan, Ye, Siqi Zhang, Kang Duan, Yong Xu, Kaixuan Guo, Cheng Chen, Somboon Chaemchuen, Dongfeng Cao, and Francis Verpoort. 2024. "Multifunctional Biomass-Based Ionic Liquids/CuCl-Catalyzed CO2-Promoted Hydration of Propargylic Alcohols: A Green Synthesis of α-Hydroxy Ketones" International Journal of Molecular Sciences 25, no. 3: 1937. https://doi.org/10.3390/ijms25031937

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