A dual-functional catalyst for selective dicarbamate synthesis via oxidative carbonylation: enhanced methoxylation for suppressing urea polymer formation

Seulgi Han ab, Young-Woo You ab, Kwanyong Jeong a, Mintaek Im a, Jung-Ae Lim a, Soo Min Kim a, Jin Hee Lee *ab and Ji Hoon Park *ab
aCO2&Energy Research Center, Korea Research Institute of Chemical Technology (KRICT), Daejeon 34114, Republic of Korea. E-mail: leejh@krict.re.kr; jihpark@krict.re.kr
bAdvanced Materials and Chemical Engineering Technology, University of Science and Technology (UST), Daejeon 34113, Republic of Korea

Received 1st March 2024 , Accepted 19th March 2024

First published on 20th March 2024


Abstract

High-yield dicarbamate synthesis via oxidative carbonylation has been challenging due to substantial urea polymer formation. A Pd/CeO2 catalyst was successfully synthesized dicarbamates while suppressing urea polymer formation. The superior methoxylation reactivity of Pd/CeO2, which originated from Pd–CeO2 synergism, facilitated the conversion of the urea species, thereby preventing urea accumulation.


With the growing concern over the climate crisis and environmental issues, the demand for low-carbon and eco-friendly production is continuously increasing. Polyurethane (PUR), an important petrochemical with wide applications such as packaging, furniture, coatings, adhesives, and thermoplastics,1,2 is also subject to these environmental demands. PUR is traditionally synthesized by the reaction of polyols and diisocyanates. The green synthesis of polyols is well-established, achieving a significant reduction in the carbon footprint by up to 3 kg per kg of CO2 incorporated,3 whereas diisocyanate synthesis still relies on petroleum-derived chemicals with high carbon footprints and extremely toxic phosgene.4,5 Therefore, it is essential to develop green diisocyanate synthesis processes that employ low-carbon materials and avoid the use of phosgene for eco-friendly PUR production.

Non-phosgene diisocyanates can be practically obtained through the thermal decomposition of dicarbamates, indicating that the efficient and sustainable synthesis of dicarbamates is important.6,7 Dicarbamates can be synthesized by reacting nitro- or amino compounds with carbonyl sources, such as dimethyl carbonate (DMC), CO2, and CO.8–12 The synthesis route using DMC is advantageous due to its mild reaction conditions and high carbamate yields. However, this method requires additional processes for DMC synthesis, such as transesterification of cyclic carbonates with methanol and oxidative carbonylation of methanol.7,13 The direct synthesis of dicarbamates from CO2 is highly desirable and promising, given that CO2 is abundant and nontoxic. However, due to the high thermodynamic stability and low reaction kinetics of CO2, high operating temperatures and pressures are required, and more than equivalent quantities of environmentally and industrially unfavored substances such as organic halides and metal alkoxides would be necessary.14–21 Numerous efforts have been made to overcome these drawbacks of the CO2 route, including the use of dehydrating agents to shift the thermodynamic equilibrium18 and adoption of alcohol as a more sustainable coupling reagent.20 However, these approaches still require complex reaction systems or excess reagents.

The oxidative carbonylation (OC) route, which uses CO as a carbonyl source, presents a more industrially viable method for dicarbamate synthesis. The advantages of the OC route include: (1) high atom efficiency by incorporating two C1 building blocks (CO and methanol); (2) relatively high yields, which are beneficial for practical applications; and (3) CO2 utilization, as CO can be readily produced from CO2 by reactions such as reforming, hydrogenation, and the reverse Boudouard reaction (CO2 + C → 2CO).22 Recently, as part of the Carbon2Polymer project in Germany, dicarbamate synthesis via oxidative carbonylation was conducted.4,6,23 In this process, industrially relevant aromatic dicarbamates, toluene-2,4-dicarbamate (TDC) and methylene diphenyl-4,4′-dicarbamate (MDC), were successfully synthesized with palladium-supported heterogeneous catalysts via OC. Moreover, the feasibility of processes for the diisocyanate production via the OC route was demonstrated through a basic process design.4

Oxidative carbonylation typically provides fairly high carbamate yields. 79.5% selectivity of methyl N-phenyl carbamate (MPC) was achieved on a Pd/C catalyst system with 80% conversion of aniline.24 A polymer-supported palladium–manganese bimetallic catalyst exhibited a higher yield of ethyl N-phenyl carbamate (90%).25 This superior performance of oxidative carbonylation has encouraged various studies in terms of catalysts, reaction conditions, and additives.24–28 However, most of the research on oxidative carbonylation has been devoted to monocarbamate synthesis, despite dicarbamates, which have two functional groups, being essential in polymerization reactions for PUR.24,29–32Only a few studies have dealt with dicarbamate synthesis via OC.33–36 Leitner's group recently reported a novel catalytic system with heterogeneous Pd catalysts for dicarbamate synthesis. This system yielded 44% TDC and 24% MDC.6,37 They elucidated the reaction network by thoroughly analyzing products and investigated the effect of the support on dicarbamate yields. Another key finding of this study was the identification of yield limitations not observed in the monocarbamate synthesis. They suggested that a major cause of the limited dicarbamate yields was the formation of a considerable amount of solid precipitates, mainly consisting of urea derivatives.

 
image file: d4gc01045k-u1.tif(1)
 
image file: d4gc01045k-u2.tif(2)

The formation of solid precipitates (urea polymer) during dicarbamate synthesis can be explained by the OC reaction pathway. The dicarbamate synthesis consists of diamine carbonylation, which forms urea intermediates (eqn (1)), and subsequent urea methoxylation to convert the urea intermediates to carbamates (eqn (2)).38–40

This indicates that when the methoxylation rate is slower than the carbonylation rate, the urea intermediates continue to grow and produce high-molecular-weight and unreactive urea polymer (Scheme 1). Therefore, developing a catalytic system that balances the rates of carbonylation and methoxylation by accelerating the methoxylation is crucial in reducing the formation of urea polymer.


image file: d4gc01045k-s1.tif
Scheme 1 Dicarbamate synthesis via oxidative carbonylation of diamines.

Herein, we identified catalytic and reaction characteristics necessary to balance the carbonylation and methoxylation rates in dicarbamate synthesis via OC. By conducting carbamate synthesis under various reaction conditions and analyzing the individual reaction rates of carbonylation and methoxylation, we found that enhancing the methoxylation rate is critical in minimizing the formation of urea polymer. It is well known that catalyst supports significantly influence the physicochemical properties of active sites and can directly participate in chemical reactions. Thus, we investigated the effects of the catalyst support on methoxylation reactivity using several supports (Al2O3, SiO2, and CeO2) that are typically used in catalytic reactions. Notably, the interaction between Pd and CeO2 increased the methoxylation reactivity by activating urea and methanol, leading to the successful synthesis of aromatic dicarbamates using the Pd/CeO2 catalyst with minimal solid precipitate formation.

We employed Pd catalysts supported on metal oxides for the oxidative carbonylation of 2,4-toluenediamine (TDA) (Fig. 1a). All the prepared catalysts were characterized using X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and transmission electron microscopy (TEM) (Fig. S1–3). TDA is completely consumed and converted over all the catalysts, but TDC yields vary considerably depending on the catalysts. Pd/CeO2 shows the highest TDC yield among the supported Pd catalysts, and the yield is even higher than that obtained using PdCl2, which is known to be an effective catalyst for the oxidative carbonylation of aniline.40–42 The higher performance of Pd/CeO2 is notable since PdCl2, a homogeneous catalyst, has many more accessible Pd sites than Pd/CeO2 (Fig. S1 and S3). This result indicates that the number of exposed Pd sites is not the only factor that results in higher TDC productivity. Moreover, the different TDC yields depending on the supports indicate that the catalyst support plays an important role in TDC synthesis.


image file: d4gc01045k-f1.tif
Fig. 1 TDC synthesis over homo- and heterogeneous Pd catalysts using different (a) catalysts, (b) temperature and (c) pressure. Reaction conditions: Pd/CeO2 catalyst, Pd (2.8 mol%), TDA (0.03 M), NaI (30 mol%), CO/O2 = 4, pressure (10 bar), temperature 135 °C, time 5 h. Reaction conditions for (d): pressure (5 bar). Reaction conditions for Pd/CeO2(×2): Pd (5.6 mol%), NaI (60 mol%), time 10 h. The yield was determined using HPLC.

Insoluble solid matter was observed after the reaction with PdCl2. Conversely, organic solid matter was rarely observed for Pd/CeO2, as confirmed by thermogravimetric analysis (TGA) of the spent catalysts (Fig. 2). The weight loss in TGA measurements is due to thermal oxidation of organic solids that cannot be removed by solvent washing or drying. Pd/CeO2 showed the lowest weight loss, indicating the lowest organic solid formation, followed by Pd/Al2O3 and later Pd/SiO2. The TDC yield of Pd/CeO2 was maintained consistently even after five recycling runs (Fig. S4). However, Pd/Al2O3 and Pd/SiO2 showed an obvious decline in the TDC yields with each cycle. This was probably due to the residual organics on the catalyst surface, hindering the interaction between the active sites and reactants. As described above, this organic solid could possibly be urea polymer, which does not further react to produce TDC (Scheme 1).6 Since the urea polymer with high molecular weight forms when the amine carbonylation is faster than urea methoxylation, it is necessary to suppress urea polymer formation by controlling the reaction rates of carbonylation and/or methoxylation to obtain a high TDC yield.


image file: d4gc01045k-f2.tif
Fig. 2 TGA curves of fresh and spent Pd catalysts.

The effects of the reaction conditions on the reaction rates of carbonylation and methoxylation were investigated (Fig. 1b–d). The TDC yields for different reaction temperatures are shown in Fig. 1b. The yield increases in the order of 110 °C < 160 °C < 135 °C. TGA results showed that this order is closely related to the amount of urea polymer formation (Fig. S5). The amount of urea polymer, as determined by weight loss in TGA, was 22.6, 5.5, and 8.9% after the reactions at 110, 135, and 160 °C, respectively. The reaction temperature showing the highest TDC yield exhibited the lowest weight loss and vice versa. The carbonylation reaction for urea synthesis from aniline is known to take place at around 100 °C.39,43,44 However, MPC synthesis (aniline carbonylation + 1,3-diphenylurea (DPU) methoxylation) requires a much higher temperature of 150 °C,29,31 indicating a higher reaction temperature for methoxylation than carbonylation. We can infer that the cause of the largest amount of urea polymer formation and the lowest TDC yield at 110 °C is the slower methoxylation rate relative to the carbonylation rate at this reaction temperature. Considering that the amount of urea polymer increases again at 160 °C, TDC synthesis is also not favourable at an excessively high temperature, and it appears that amine carbonylation and urea methoxylation rates are well balanced at 135 °C in our catalytic system.

The TDC yield and reaction pressure show an inverse relationship (Fig. 1c). Namely, the TDC yield increases as the reaction pressure decreases. In general, high pressure accelerates the overall reaction rate of carbamate synthesis, which is attributed to the increased CO and O2 solubility.45,46 However, this is also related to the respective reaction rates of amine carbonylation and urea methoxylation. The reaction pressure of 5 bar might be high enough for amine carbonylation to proceed and low enough to hinder urea polymer formation. Conclusively, we proved the significant impact of the catalysts and reaction conditions on urea polymer formation and TDC yield. This finding suggests that adjusting catalytic properties and the reaction conditions is a viable strategy to achieve a balance between the carbonylation and methoxylation rates.

Finally, we compared the TDC yields over PdCl2 and Pd/CeO2 under the optimized conditions (Fig. 1d). The highest TDC yield of 53% was obtained over Pd/CeO2. In contrast, PdCl2 shows only a 27% TDC yield under the same reaction conditions. Since the conversion was 100% on both catalysts, the difference in the TDC yield is due to a difference in methoxylation ability. In addition, when the amounts of catalysts and reaction time were increased twofold, the TDC yield was increased by up to 80%. As shown above, the urea polymer formation is a decisive hurdle in TDC synthesis, and the different reaction rates between carbonylation and methoxylation are the most important factor in this regard.

We thus carried out MPC synthesis from aniline to reveal the rate difference between amine carbonylation and urea methoxylation over Pd/CeO2 and PdCl2 (Fig. 3). Aniline, which has a single amine functional group, was applied as a substrate. The MPC synthesis from aniline via oxidative carbonylation proceeds via two sequential reactions: carbonylation, which converts aniline to DPU, and methoxylation, which converts DPU to MPC, which are identical processes to TDC synthesis (Scheme 1). In MPC synthesis over Pd/CeO2, aniline conversion and MPC yield rapidly increased from 20 to 120 min, showing 96% conversion and 91% yield. The similar slopes of the aniline conversion and MPC yield indicate that the carbonylation and methoxylation rates are similar on Pd/CeO2. The yield of the intermediate DPU increased initially, but decreased after 40 min, resulting in a high yield of MPC. On the contrary, different reaction rates between carbonylation and methoxylation were confirmed on the PdCl2 catalyst. The slope of the MPC yield is lower than that of the aniline conversion, indicating a slow methoxylation rate. As a result, the yield of the intermediate DPU steadily increased up to 60 min and further decreased much more slowly compared to that of Pd/CeO2. These experiments provide evidence that the high MPC yield of Pd/CeO2 is essentially attributable to the fast methoxylation rate. The fast methoxylation rate of Pd/CeO2 favorably affected the TDC yield, as confirmed in Fig. 1.


image file: d4gc01045k-f3.tif
Fig. 3 Time course of MPC synthesis over Pd/CeO2 and PdCl2. Reaction conditions: aniline (0.06 M), Pd (1.4 mol%), NaI (30 mol%), temperature 135 °C, pressure 5 bar. Conversion and yield were determined using GC-MS.

To elucidate the effects of the Pd catalyst supports on the methoxylation rates, DPU methoxylation was tested using Pd/CeO2, Pd/Al2O3, and Pd/SiO2 (Table 1). As described above, this reaction can reflect the methoxylation reaction in the TDC synthesis. The MPC yields follow the order Pd/CeO2 > Pd/Al2O3 > Pd SiO2, which correlates with the TDC yields (Fig. 1a). These results demonstrate that the methoxylation of the urea intermediate is the key step in the overall dicarbamate synthesis, and the high activity of Pd/CeO2 in TDC synthesis is caused by the accelerated methoxylation. The roles of the individual catalyst components for the carbonylation and methoxylation were investigated. The carbonylation only proceeds in the presence of Pd and NaI. The absence of either one results in no conversion of amine (Table S1), indicating that Pd–I acts as the active site for the carbonylation reaction. This finding is highly consistent with previous research.24,47 Analysis of entries 4–6 in Table 1 reveals that using CeO2 by itself or adding NaI does not lead to any improvement in the methoxylation reaction. However, a significant increase in carbamate yield is observed for the Pd/CeO2 catalyst. This indicates that the synergistic interaction between Pd and CeO2 has a positive impact on the methoxylation rate.

Table 1 DPU methoxylation using different catalysts and NaI

image file: d4gc01045k-u3.tif

Entry Catalyst NaI (mol%) MPC yield (%)
1 Pd/CeO2 95
2 Pd/Al2O3 81
3 Pd/SiO2 45
4 Pd/CeO2 30 94
5 CeO2 30 66
6 67


An XPS analysis was performed to identify the origin of the high methoxylation rate of the Pd/CeO2 catalyst. Pd 3d XPS spectra consisting of Pd 3d5/2 and Pd3/2 are presented in Fig. 4a, and the Pd2+/Pd0 ratio was calculated through peak fitting deconvolution. Pd/CeO2 exhibits a higher Pd2+ ratio (42%) than Pd/SiO2 (32%) and Pd/Al2O3 (31%). Fig. 4b shows the Ce 3d spectra of CeO2 and Pd/CeO2. The oxidation state of CeO2 was identified as Ce4+ (filled in yellow) with a small portion of Ce3+ (filled in purple). Conversely, Pd/CeO2 shows obviously increased Ce3+ peaks compared to CeO2. This high ratio of Pd2+ to Ce3+ in the Pd/CeO2 catalyst is due to the substitution of Ce4+ by Pd2+, as reported in previous research.48 The effect of Pd2+ and Ce3+ on the methoxylation rates can be inferred from the literature. According to Kostic's report, Pd2+ efficiently catalyzes the alcoholysis of urea, which is a similar process to the methoxylation of the urea intermediate.49 The activation of the carbonyl moiety in urea by Pd2+ facilitates the nucleophilic attack of methanol, resulting in an increased methoxylation rate, which ultimately leads to a high yield of TDC. Somorjai et al. reported the effect of Ce3+ on methanol decomposition. They demonstrated that Ce3+ sites with a neighboring oxygen vacancy (Ce3+–Ov–Ce3+) enhance the adsorption capacity of methanol and stabilize adsorbed methoxy species well. The reactivity of Ce3+ with methanol is beneficial to the methoxylation reaction in TDC synthesis. Based on these reports, it can be concluded that the superior methoxylation activity of Pd/CeO2 originates from its high ratios of Pd2+ and Ce3+. This electronic structure of Pd/CeO2 remained stable even after the reaction (Fig. S2), contributing to the consistent catalyst reactivity (Fig. S4).


image file: d4gc01045k-f4.tif
Fig. 4 (a) Pd 3d XPS spectra of reduced Pd/CeO2, Pd/Al2O3 and Pd/SiO2. (b) Ce 3d XPS spectra of reduced CeO2 and Pd/CeO2. (c) FT-IR spectra of the catalysts after sequential exposure to methanol/N2 and N2 flow at 50 °C for 1 h.

In situ diffuse reflectance infrared transform spectroscopy (DRIFTS) experiments were conducted with a methanol flow to elucidate the difference in the methoxylation rate for the different catalysts. Methanol/N2 gas was introduced into the catalyst for 30 min, followed by N2 purging for 1 h at 50 °C (the detailed experimental procedure is described in the ESI). Strong gaseous methanol peaks, which make it difficult to observe chemisorbed species, were mainly observed under the methanol/N2 flow (Fig. S6). After N2 purging to remove the gaseous methanol and physisorbed species, the chemisorbed species were observed. Fig. 4c shows Fourier transform infrared (FT-IR) spectra in the range of the C–O vibration of the methoxy species after N2 purging. Three major peaks are observed at 1098, 1043, and 1022 cm−1, corresponding to monodentate (1098 cm−1) and multidentate (1043 and 1022 cm−1) bridging methoxy species.50–52

As can be seen in results, Pd/CeO2 shows much higher intensity than the other catalysts. Relatively small peaks are observed for Pd/Al2O3. Pd/SiO2 exhibits no characteristic adsorption peak. These results indicate that the abundance of the surface methoxy species follows the order Pd/CeO2 > Pd/Al2O3 > Pd/SiO2. The large amount of surface methoxy species on Pd/CeO2 promotes urea methoxylation (refer to Table 1), thereby increasing the TDC yield (refer to Fig. 1a). We propose a reaction mechanism for dicarbamate synthesis via OC, based on the findings of this study and relevant literature. Initially, Pd metal is activated by NaI to form a Pd–I bond, then reacts with CO and amine, leading to the formation of a carbamoyl species. Subsequently, the carbamoyl species reacts with another amine to produce the urea intermediate. This urea intermediate then undergoes reaction with methanol to yield dicarbamate and diamine. The surface hydrogen generated by the deprotonation of amine is removed by oxygen, leading to the regeneration of the catalyst surface. The proposed reaction mechanism is illustrated in Scheme S1.

We also applied our catalytic system to other diamine and alcohol substrates (Scheme 2). Pd/CeO2 successfully synthesizes MDC, which is a precursor of MDI, occupying more than 40% of the PU market, via the oxidative carbonylation of 4,4′-methylenediphenyl diamine (MDA) (1). In contrast to that of aromatic amines, the synthesis of aliphatic amines did not proceed efficiently. A 1,6-hexanedicarbamate (2) yield of only 32% was observed, and isophorone dicarbamate (3) was not synthesized at all. Although the yield of aliphatic dicarbamates was low, the amine species was completely converted. This result indicates that the low aliphatic dicarbamate yield is most likely due to the high nucleophilicity of the aliphatic amine, which promotes urea polymer formation. The chain length of the alcohol greatly affected the dicarbamate yield (4–6). As the alkyl chain becomes longer, the corresponding dicarbamate yields are lower. It appears that short chain alcohols are better in terms of acting as a nucleophile for the substitution reaction. In addition, it is supposed that the longer the alcohol chain is, the lower the urea intermediate solubility is, which may inhibit alcoholysis, making it difficult to convert to carbamate.53


image file: d4gc01045k-s2.tif
Scheme 2 Oxidative carbonylation using various amines and alcohols.

In this study, we attempted to synthesize dicarbamates in high yields by alleviating the formation of urea polymer and to elucidate the relationship between the dicarbamate yield and urea polymer formation. Pd/CeO2 showed a higher yield of TDC than the other catalysts, including homogeneous PdCl2. A fairly high TDC yield of 80% was achieved over Pd/CeO2 under mild reaction conditions (135 °C and 5 bar). The catalysts with high TDC yield showed low urea polymer formation and vice versa. Given that urea polymer can be formed when the methoxylation rate is slower than the carbonylation rate, the small amount of urea polymer over Pd/CeO2 was caused by the fast methoxylation rate. Indeed, the superior methoxylation reactivity of Pd/CeO2 was confirmed in the MPC synthesis and DPU methoxylation reaction. XPS and DRIFT experiments were conducted to investigate the origin of the superior methoxylation of Pd/CeO2. XPS results indicated that Pd/CeO2 had high ratios of Pd2+ and Ce3+, which are known to be beneficial for the activation of urea and methanol, respectively. In the DRIFT experiment under a methanol flow, a much higher amount of surface methoxy species was observed in Pd/CeO2 compared to the other catalysts. The abundant methoxy species on the catalyst surface can promote the methoxylation rate, which in turn leads to a high yield of dicarbamates. We believe that these findings will encourage the development of efficient catalytic systems for eco-friendly non-phosgene isocyanate production.

Author contributions

S. H: data curation, formal analysis, investigation, methodology, writing – original draft. Y-W. Y: formal analysis, investigation, resources, writing – review & editing. K. J: investigation, resources. M. I: investigation, resources. J-A. L: investigation, resources. J. H. L: methodology, funding acquisition, supervision, validation, writing – review & editing. J. H. P: conceptualization, methodology, project administration, funding acquisition, supervision, validation, writing – review & editing.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by Institutional Research Program of KRICT (KK2411-20) and Technology Innovation Program (20225A10100040) funded by the Ministry of Trade, Industry & Energy.

References

  1. D. K. Chattopadhyay and K. V. S. N. Raju, Prog. Polym. Sci., 2007, 32, 352–418 CrossRef CAS.
  2. H. M. C. C. Somarathna, S. N. Raman, D. Mohotti, A. A. Mutalib and K. H. Badri, Constr. Build. Mater., 2018, 190, 995–1014 CrossRef CAS.
  3. N. von der Assen and A. Bardow, Green Chem., 2014, 16, 3272–3280 RSC.
  4. T. Kaiser, A. Rathgeb, C. Gertig, A. Bardow, K. Leonhard and A. Jupke, Chem. Ing. Tech., 2018, 90, 1497–1503 CrossRef CAS.
  5. S. A. Cucinell and E. Arsenal, Arch. Environ. Health, 1974, 28, 272–275 CrossRef CAS PubMed.
  6. C. Hussong, J. Langanke and W. Leitner, Chem. Ing. Tech., 2020, 92, 1482–1488 CrossRef CAS.
  7. P. Wang, S. Liu and Y. Deng, Chin. J. Chem., 2017, 35, 821–835 CrossRef CAS.
  8. O. Kreye, H. Mutlu and M. A. R. Meier, Green Chem., 2013, 15, 1431–1455 RSC.
  9. Y. Cao, Y. Chi, A. Muhammad, P. He, L. Wang and H. Li, Chin. J. Chem. Eng., 2019, 27, 549–555 CrossRef CAS.
  10. Y. Dai, Y. Wang, J. Yao, Q. Wang, L. Liu, W. Chu and G. Wang, Catal. Lett., 2008, 123, 307–316 CrossRef CAS.
  11. R. Juárez, A. Corma and H. García, Top. Catal., 2009, 52, 1688–1695 CrossRef.
  12. D.-L. Sun, J.-Y. Luo, R.-Y. Wen, J.-R. Deng and Z.-S. Chao, J. Hazard. Mater., 2014, 266, 167–173 CrossRef CAS PubMed.
  13. H. Wang, B. Lu, X. Wang, J. Zhang and Q. Cai, Fuel Process. Technol., 2009, 90, 1198–1201 CrossRef CAS.
  14. Q. Zhang, H.-Y. Yuan, N. Fukaya, H. Yasuda and J.-C. Choi, ChemSusChem, 2017, 10, 1501–1508 CrossRef CAS PubMed.
  15. D. Riemer, P. Hirapara and S. Das, ChemSusChem, 2016, 9, 1916–1920 CrossRef CAS PubMed.
  16. Q. Zhang, H.-Y. Yuan, N. Fukaya and J.-C. Choi, ACS Sustainable Chem. Eng., 2018, 6, 6675–6681 CrossRef CAS.
  17. K. A. Alferov, Z. Fu, S. Ye, D. Han, S. Wang, M. Xiao and Y. Meng, ACS Sustainable Chem. Eng., 2019, 7, 10708–10715 CrossRef CAS.
  18. Y. Gu, A. Miura, M. Tamura, Y. Nakagawa and K. Tomishige, ACS Sustainable Chem. Eng., 2019, 7, 16795–16802 CrossRef CAS.
  19. Q. Zhang, H.-Y. Yuan, N. Fukaya, H. Yasuda and J.-C. Choi, Green Chem., 2017, 19, 5614–5624 RSC.
  20. Q. Zhang, H.-Y. Yuan, X.-T. Lin, N. Fukaya, T. Fujitani, K. Sato and J.-C. Choi, Green Chem., 2020, 22, 4231–4239 RSC.
  21. M. Abla, J.-C. Choi and T. Sakakura, Green Chem., 2004, 6, 524–525 RSC.
  22. P. Lahijani, Z. A. Zainal, M. Mohammadi and A. R. Mohamed, Renewable Sustainable Energy Rev., 2015, 41, 615–632 CrossRef CAS.
  23. W. Leitner, G. Franciò, M. Scott, C. Westhues, J. Langanke, M. Lansing, C. Hussong and E. Erdkamp, Chem. Ing. Tech., 2018, 90, 1504–1512 CrossRef CAS.
  24. S. P. Gupte and R. V. Chaudhari, J. Catal., 1988, 114, 246–258 CrossRef CAS.
  25. B. Wan, S. Liao and D. Yu, Appl. Catal., A, 1999, 183, 81–84 CrossRef CAS.
  26. S. Fukuoka, M. Chono and M. Kohno, J. Chem. Soc., Chem. Commun., 1984, 399–400 RSC.
  27. H. Alper and F. W. Hartstock, J. Chem. Soc., Chem. Commun., 1985, 1141–1142 RSC.
  28. B. Chen and S. S. C. Chuang, J. Mol. Catal. A: Chem., 2003, 195, 37–45 CrossRef CAS.
  29. S. Kanagasabapathy, A. Thangaraj, S. P. Gupte and R. V. Chaudhari, Catal. Lett., 1994, 25, 361–364 CrossRef CAS.
  30. F. Ferretti, E. Barraco, C. Gatti, D. R. Ramadan and F. Ragaini, J. Catal., 2019, 369, 257–266 CrossRef CAS.
  31. F.-M. Mei, L.-J. Chen and G.-X. Li, Appl. Organomet. Chem., 2010, 24, 86–91 CrossRef CAS.
  32. P. Toochinda and S. S. C. Chuang, Ind. Eng. Chem. Res., 2004, 43, 1192–1199 CrossRef CAS.
  33. F. Shi and Y. Deng, J. Catal., 2002, 211, 548–551 CAS.
  34. F. Shi and Y. Deng, Chem. Commun., 2001, 443–444 RSC.
  35. F. Shi, Y. Deng, T. SiMa and H. Yang, J. Catal., 2001, 203, 525–528 CrossRef CAS.
  36. V. Valli and H. Alper, Organometallics, 1995, 14, 80–82 CrossRef CAS.
  37. C. Hussong, J. Langanke and W. Leitner, Green Chem., 2020, 22, 8260–8270 RSC.
  38. F. Ragaini, Dalton Trans., 2009, 6251–6266 RSC.
  39. B. Gabriele, G. Salerno, R. Mancuso and M. Costa, J. Org. Chem., 2004, 69, 4741–4750 CrossRef CAS PubMed.
  40. A. Krogul and G. Litwinienko, J. Mol. Catal. A: Chem., 2015, 407, 204–211 CrossRef CAS.
  41. H. Alper, G. Vasapollo, F. W. Hartstock, M. Mlekuz, D. J. H. Smith and G. E. Morris, Organometallics, 1987, 6, 2391–2393 CrossRef CAS.
  42. P. Giannoccaro, J. Organomet. Chem., 1987, 336, 271–278 CrossRef CAS.
  43. S. Liu, X. Dai, H. Wang and F. Shi, Green Chem., 2019, 21, 4040–4045 RSC.
  44. S. T. Gadge, E. N. Kusumawati, K. Harada, T. Sasaki, D. Nishio-Hamane and B. M. Bhanage, J. Mol. Catal. A: Chem., 2015, 400, 170–178 CrossRef CAS.
  45. F. Wu, Q. Zhao, L. Tao, D. Danaci, P. Xiao, F. A. Hasan and P. A. Webley, J. Chem. Eng. Data, 2019, 64, 5609–5621 CrossRef CAS.
  46. K. Fischer and M. Wilken, J. Chem. Thermodyn., 2001, 33, 1285–1308 CrossRef CAS.
  47. A. A. Kelkar, D. S. Kolhe, S. Kanagasabapathy and R. V. Chaudhari, Ind. Eng. Chem. Res., 1992, 31, 172–176 CrossRef CAS.
  48. Y.-Q. Su, J.-X. Liu, I. A. W. Filot, L. Zhang and E. J. M. Hensen, ACS Catal., 2018, 8, 6552–6559 CrossRef CAS PubMed.
  49. N. V. Kaminskaia and N. M. Kostić, Inorg. Chem., 1998, 37, 4302–4312 CrossRef CAS PubMed.
  50. Z. Qi, L. Chen, S. Zhang, J. Su and G. A. Somorjai, J. Am. Chem. Soc., 2021, 143, 60–64 CrossRef CAS PubMed.
  51. E. Finocchio, M. Daturi, C. Binet, J. C. Lavalley and G. Blanchard, Catal. Today, 1999, 52, 53–63 CrossRef CAS.
  52. A. Badri, C. Binet and J.-C. Lavalley, J. Chem. Soc., Faraday Trans., 1997, 93, 1159–1168 RSC.
  53. F.-M. Lee and L. E. Lahti, J. Chem. Eng. Data, 1972, 17, 304–306 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4gc01045k

This journal is © The Royal Society of Chemistry 2024