Next Article in Journal
Highlights on the Catalytic Properties of Polyoxometalate-Intercalated Layered Double Hydroxides: A Review
Next Article in Special Issue
Palladium PEPPSI-IPr Complex Supported on a Calix[8]arene: A New Catalyst for Efficient Suzuki–Miyaura Coupling of Aryl Chlorides
Previous Article in Journal
Biochemical and Structural Characterization of Cross-Linked Enzyme Aggregates (CLEAs) of Organic Solvent Tolerant Protease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rhodium-Catalyzed Aqueous Biphasic Olefin Hydroformylation Promoted by Amphiphilic Cyclodextrins

by
Aurélien Cocq
1,
Hervé Bricout
1,
Florence Djedaïni-Pilard
2,
Sébastien Tilloy
1 and
Eric Monflier
1,*
1
Univ. Artois, CNRS, Centrale Lille, ENSCL, Univ. Lille, Unité de Catalyse et Chimie du Solide (UCCS), UMR 8181, rue Jean Souvraz, F-62307 Lens, France
2
Laboratoire de Glycochimie des Antimicrobiens et des Agroressources, LG2A UMR 7378 CNRS, Université de Picardie Jules Verne, 33 rue Saint-Leu, 80039 Amiens, France
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(1), 56; https://doi.org/10.3390/catal10010056
Submission received: 5 December 2019 / Revised: 18 December 2019 / Accepted: 20 December 2019 / Published: 1 January 2020
(This article belongs to the Special Issue Supramolecules for Catalysis)

Abstract

:
Hydroformylation is an industrial process that allows for the production of aldehydes from alkenes using transition metals. The reaction can be carried out in water, and the catalyst may be recycled at the end of the reaction. The industrial application of rhodium-catalyzed aqueous hydroformylation has been demonstrated for smaller olefins (propene and butene). Unfortunately, larger olefins are weakly soluble in water, which results in very low catalytic activity. In an attempt to counteract this, we investigated the use of amphiphilic oleic succinyl-cyclodextrins (OS-CDs) synthesized from oleic acid derivatives and maleic anhydride. OS-CDs were found to increase the catalytic activity of rhodium during the hydroformylation of water-insoluble olefins, such as 1-decene and 1-hexadecene, by promoting mass transfer. Recyclability of the catalytic system was also evaluated in the presence of these cyclodextrins.

Graphical Abstract

1. Introduction

Cyclodextrins (CDs) are cyclic oligosaccharides consisting of six (α-CD), seven (β-CD), or eight (γ-CD) α-D-glucopyranose units [1,2]. They form conical cylinders with an outer hydrophilic surface and inner hydrophobic surface. The inner surface forms a cavity and allows for the formation of an inclusion complex when it binds a hydrophobic substrate. To vary their size and shape, native CDs can be modified by substituting their hydroxyl groups with various other functional groups. These are referred to as mono- or polysubstituted CDs. Due to their ability to form inclusion complexes with a wide range of compounds, native and modified CDs have applications in various fields. These include analytical chemistry, agriculture, pharmaceuticals, foods, toilet products, and catalysis [3]. In the field of aqueous biphasic organometallic catalysis, CDs have mainly been employed as platforms for the production of water-soluble ligands using transition metals [4,5,6,7,8,9,10,11,12,13]. They have also been used as mass-transfer agents in order to increase the solubility of hydrophobic substrates in water [14,15]. For example, the use of CDs in rhodium-catalyzed hydroformylation in an aqueous/organic two-phase system was previously described [16,17,18,19,20]. Increases in catalytic potential were especially high for hydrophobic substrates, which shows the beneficial effect of CDs on mass transfer at the aqueous–organic interface. However, while the effect of CDs was experimentally observed in 1995 [21], their exact role was only clearly established about 10 years later. Indeed, a molecular-dynamics study by Sieffert and Wipff described the interactions occurring between olefins, the catalytic system, and CDs at the aqueous–organic interface [22,23]. The study was performed in accordance with surface-tension measurements [24]. It was shown that the mass-transfer properties of CDs (and consequently the catalytic activity) were linked to surface excess, i.e., the number of CDs adsorbed per m2 of interface. Modified CDs (mainly randomly methylated β-CD (RAME-β-CD)) increased the hydroformylation rate of olefins possessing eight to twelve carbons. However, CD efficiency was drastically reduced for larger olefins (>12 carbons) [25]. Indeed, the shape and size of the CD cavity did not allow for the binding of larger olefins, leading to lower catalytic activity. The construction of more sophisticated CDs, such as CD dimers or polymers, is therefore required to accommodate larger olefins with multivalent interactions [15]. For this reason, we evaluated the behavior of amphiphilic oleic succinyl-CDs (OS-CDs) synthesized from oleic acid derivatives and maleic anhydride in an aqueous hydroformylation reaction (Scheme 1). These CDs are easily synthetized, highly soluble in water, have low surface tension, and have low critical-aggregation concentration (CAC). The formation of aggregates could be useful in the solubilization of larger olefins in water. Surfactants have been used to facilitate hydroformylation reactions between water-soluble catalysts and larger olefins [26,27,28,29,30]. In this report, we showed that various OS-CDs were efficient in promoting the hydroformylation of water-insoluble olefins such as 1-decene and 1-hexadecene. The recyclability of the catalytic system was also evaluated in the presence of these CDs.

2. Results and Discussion

OS-CDs synthesis was conducted in two steps as previously reported by our group [31,32]. The first step involved the alkenylation of maleic anhydride by methyl oleate and oleic acid, leading to the formation of maleinised methyl oleate or maleinised oleic acid, respectively. This step yielded a mixture of four regioisomers, with each regioisomer consisting of eight stereoisomers. The mixture is represented hereafter as a single compound (Scheme 1, compounds 1 and 2). The second step involved the grafting of maleinised methyl oleate or maleinised oleic acid onto native CDs (α-, β-, or γ-CD), RAME-β-CD, and hydroxypropylated-β-CD (HP-β-CD). This was followed by the neutralization of carboxyl groups by sodium hydroxide. The obtained products were named CD-(OSG–Z)DS, where CD corresponded to the nature of the cyclodextrin, OSG stood for Oleic Succinic Graft, Z indicated a methyl group or sodium atom (according to whether the reactant was methyl oleate or oleic acid), and DS was the degree of substitution.
Some physicochemical properties of OS-CDs are listed in Table 1.
The effect of these OS-CDs was studied in the rhodium-catalyzed hydroformylation of two olefins, 1-decene and 1-hexadecene. Catalysis experiments were conducted in an aqueous biphasic medium using Rh(acac)(CO)2 as catalyst precursor and trisulfonated triphenylphosphine (TPPTS) as a ligand (Table 2). The reaction was carried out in an autoclave with a mechanical stirrer in order to elicit the maximum effect on mass transfer between the two layers. The pressure of syngas (CO:H2 = 1:1) was 50 bar, the stirring rate was 1500 rpm, and the temperature was 80 °C. The first experiment was performed without a promotor, and therefore produced very little product in 6 h due to the very low water solubility of 1-decene and 1-hexadecene (Table 2, entries 1 and 9). The reaction was limited by mass transfer between the two layers. Among the modified CDs presented in Table 1, β-CD-(OSG–Me)1.6 was chosen as the first promoter to be tested because it exhibited the best compromise between ease of synthesis and physicochemical properties. In addition, β-CD molecules are most appropriate for the functionalization of larger olefins [16]. Using an equivalent amount of β-CD-(OSG–Me)1.6 and Rh, 100% of 1-decene and 65% of 1-hexadecene were converted in 4 and 6 h, respectively (Table 2, entries 2 and 10). Logically, the conversion rate significantly decreased with alkyl chain length. An identical catalytic experiment was also performed for 1-decene, and we found that conversion percentages at 0.5 and 2 h was 85% and 93%, respectively (Table 2, entries 4 and 3). The chemoselectivity of aldehydes was 61% and 52% (entries 2 and 10), and the main side products were isomeric olefins. With regard to regioselectivity, the linear/branched aldehyde ratios were 2.3 and 2.8 (entries 2 and 10). These results were compared to those obtained for β-CDs and RAME-β-CD (at equivalent levels of Rh), which are classically used CDs in aqueous organometallic catalysis (Table 2, entries 5, 6, 11, and 12). For β-CD and RAME-β-CD, conversion levels of 1-decene were 6% and 21% after 6 h (Table 2, entries 5 and 6), and 3% and 9% for 1-hexadecene (Table 2, entries 11 and 12), respectively. The superior conversion rates obtained using RAME-β-CD over β-CD were previously described [16]. The chemoselectivity and regioselectivity of RAME-β-CD were 90 and 2.1, respectively. In order to evaluate the role of each moiety, other control experiments were performed in the presence of 1.6 equivalents (with respect to rhodium) of oleic moiety (4) or the combination of 1 + 1.6 equivalents of β-CD + (4). The term “1.6 equivalents” corresponds to the mean number of grafts on CDs, i.e., the DS of β-CD-(OSG–Me)1.6. Conversion percentages in the presence of compound 4 and β-CD + compound 4 were increased to 20% and 45% for 1-decene, and to 10% and 23% for 1-hexadecene, respectively. However, these values were lower than those obtained for β-CD-(OSG–Me)1.6 (Table 2, entries 7–8 and 13–14 compared to 2 and 10). In both cases, chemoselectivity was 53% and regioselectivity between 2.5 and 2.7. Regardless of the used olefin (1-decene or 1-hexadecene), conversion efficiency was higher in the presence of β-CD-(OSG–Me)1.6 than β-CD, RAME-β-CD, compound 4, or a combination β-CD + compound 4. Nevertheless, chemoselectivity was lower in the presence of β-CD-(OSG–Me)1.6. This was comparable to systems without CDs (Table 2, entries 1 and 9). The formation of aggregates with β-CD-(OSG–Me)1.6 did not prevent an isomerization reaction. In contrast, RAME-β-CD formed inclusion complexes with the substrate, which prevented isomerization of the double bond [33].
After proving that β-CD-(OSG–Me)1.6 was able to facilitate the hydroformylation of 1-decene and 1-hexadecene, other control experiments were performed with 1-hexadecene. This substrate was more challenging to functionalize in an aqueous medium.
Given that the oleic moiety of β-CD-(OSG–Me)1.6 contains a carbon–carbon double bond that can be hydroformylated, the structure of this OS-CD was analyzed after applying hydroformylation conditions without an added substrate (experiment conditions of Table 2, 1 equivalent of OS-CD per rhodium atom, no 1-decene or 1-hexadecene, and reaction time = 6 h). The aqueous phase was recovered and analyzed using matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS). By the end of this catalytic test, 16% of carbon–carbon double bonds in CD grafts were hydroformylated (as proven by the appearance of a new peak at 1581, i.e., β-CD-(OSG–Me)1 + H + CHO + Na+ on the MALDI-MS spectrum) (Figure 1). Among the mixture of CDs, only CDs bearing one graft (β-CD-(OSG–Me)1) were hydroformylated. This result may have been due to greater accessibility to the C–C double bond in the organometallic complex. Thus, in CDs bearing two, three, or four grafts, hydroformylation of unsaturated grafts rarely took place. The same MALDI-MS analysis was performed after Experiment 10 (Table 2), which led to the hydroformylation of 65% of 1-hexadecene in 6 h. In this experiment, only 6% of carbon–carbon double bonds in CD grafts were hydroformylated. Therefore, the presence of the olefin substrate decreased the hydroformylation of β-CD-(OSG–Me)1.6.
To confirm if the presence of a carbon–carbon double bond in the graft was detrimental to catalytic activity, a catalytic test was performed using β-CD-bearing hydrogenated arms (Scheme 1, β-CD-(H2-OSG–Me)1.6). To synthesize β-CD-(H2–OSG–Me)1.6, compound 1 was hydrogenated to give compound 3, which was then grafted onto β-CD. In the presence of β-CD-(H2–OSG–Me)1.6, the conversion percentage, chemoselectivity, and regioselectivity of 1-hexadecene were 68%, 58%, and 2.6, respectively. This was compared to 65%, 52%, and 2.8 in the presence of β-CD-(OSG–Me)1.6 (Table 2, entries 15 and 10). Therefore, saturation of the oleic arm did not significantly modify the course of hydroformylation.
The recyclability of the catalytic system and of β-CD-(OSG–Me)1.6 was investigated by reusing the aqueous catalytic phase obtained after Experiment 10 in successive hydroformylation runs (Table 2, entries 10, 16, and 17). Although β-CD-(OSG–Me)1.6 concentration was 380 times higher than its CAC, decantation occurred at the end of the reaction. This allowed for isolation of the aqueous layer of the reaction product. This layer was reloaded with 1-hexadecene and reused with the same experimental conditions. The aqueous layer was then used for two further runs and similar conversion percentages were obtained (Table 2, 65%, 63%, and 67% for entries 10, 16, and 17, respectively). Chemoselectivity and regioselectivity were unchanged.
Other OS-CDs presented in Table 1 were also evaluated for their capacity to act as mass-transfer promotors in the functionalization of 1-hexadecene (Table 3, entries 1–10). While the nature of CDs did not significantly influence chemoselectivity or regioselectivity (51–66 and 1.8–3.0, respectively), conversion percentages were deeply impacted—values varied from 9% to 94%. In fact, two trends emerged, one for Me-type OS-CDs possessing a methyl ester group (one sodium carboxylate group per graft, Table 3, entries 1–5), and another for Na-type OS-CDs (two sodium carboxylate groups per graft, Table 3, entries 6–10). Conversion percentages were largely higher for Me-type OS-CDs compared with Na-type OS-CDs. The tensioactive behavior of Me-type OS-CDs was more marked, given that disodium dicarboxylate derivatives have a higher CAC than that of their ester counterparts. In addition, for Na-type OS-CDs, the presence of two sodium carboxylate groups in the graft may have caused electronic repulsions between anionic groups of CDs and sulfonate groups of water-soluble phosphine. This phenomenon may have reduced catalytic activity by preventing the meeting of organometallic species and substrates. β-CD derivatives yielded a higher conversion percentage than that of α-CD derivatives. α-CD derivatives, in turn, yielded higher conversion percentages than γ-CD derivatives (Table 3, entries 1–3 and 6–8). For RAME-β-CD and HP-β-CD derivatives (Table 3, entries 4–5 and 9–10), variations in conversion percentages were significant for Me-type OS-CDs, while similar conversation percentages were obtained for Na-type OS-CDs. The best result was obtained using RAME-β-CD-(OSG–Me)0.9 (Table 3, entry 4). This CD performed 10 times better than RAME-β-CD in terms of conversion percentage, and its regioselectivity was also slightly higher, but its chemoselectivity was significantly lower. In fact, the modes of action of these two CDs were different. RAME-β-CD formed inclusion complexes with the substrate, preventing isomerization of the double bond. In the case of RAME-β-CD-(OSG–Me)0.9, the formation of aggregates facilitated binding of the substrate and catalyst but did not prevent the isomerization reaction.

3. Materials and Methods

3.1. General Remarks

Organic compounds, alkenes, and dicarbonylacetylacetonato rhodium (I) were purchased from Acros Organics, Sigma-Aldrich, and Strem Chemicals, respectively. They were purchased in their purest form and were used without further purification. The sodium salt tris (3-sodium sulfonatophenyl) phosphane (TPPTS–P(m-C6H4SO3Na)3) was synthesized using a method reported by B. Cornils et al. [34]. The purity of TPPTS was carefully controlled. In particular, 1H and 31P NMR analysis indicated that TPPTS was only sulfonated in the metaposition, and less than 1% of its oxide was present. Ultrapure water was used in all experiments (Fresenius Kabi, Paris, France). The carbon monoxide/hydrogen mixture (1:1) was used directly from cylinders (>99.9% pure, Air Liquide, Loos, France). β-CD was a generous gift from Roquette Frères (Lestrem, France). Pharmaceutical-grade RAME-β-CD (Cavasol® W7 M) was purchased from Wacker Chemie GmbH and was unmodified. All high-pressure hydroformylation experiments were carried out in a 25 mL stainless-steel autoclave supplied by Parr. NMR spectra were recorded with a Bruker DRX300 spectrometer operating at 300 MHz for 1H nuclei. CDCl3 (99.50% isotopic purity) was purchased from Euriso-Top. Mass spectra were recorded on a MALDI-TOF/TOF Bruker Daltonics Ultraflex II in positive reflectron mode with 2,5-DHB as a matrix.

3.2. General Procedure for Rhodium-Catalyzed Hydroformylation Experiments

The preparation of the catalyst and alkene was performed under nitrogen using standard Schlenk techniques. All liquid reagents were degassed by bubbling nitrogen for 15 min or using 2 freeze–pump–thaw cycles before use. Rh(acac)(CO)2 (21 µmol), TPPTS (105 µmol), and promotor (21 µmol) were dissolved in 6 mL of degassed water in an inert atmosphere. The resulting aqueous and organic phases contained alkenes (10.5 mmol), and were charged under an inert atmosphere into a 25 mL autoclave. Mechanical stirring using a multipaddle unit was then started (1500 rpm), and the reactor was heated up to 80 °C. As soon as this temperature was reached, the autoclave was pressurized with 50 bar of CO/H2 (1:1). The addition of CO/H2 was considered to be the beginning of the reaction. After the reaction, the autoclave was cooled, and the organic layer obtained after decantation was analyzed using 1H NMR.

4. Conclusions

In this study, various OS-CDs were used to facilitate the aqueous biphasic rhodium-catalyzed hydroformylation of 1-decene and 1-hexadecene. In the case of 1-hexadecene, the best catalytic activity was obtained in the presence of RAME-β-CD-(OSG–Me)0.9. Although the hydroformylation of oleic grafts in OS-CDs was observed, this did not have a detrimental effect on catalytic activity. The aqueous phase of the catalyst was recycled twice without loss of activity. The observed chemoselectivity was rather weak, but this could be explained by the formation of OS-CD aggregates that increased catalytic activity without preventing the isomerization reaction. In the case of amphiphilic CDs, the presence of a cavity did not prevent isomerization of the substrate’s double bond. These new amphiphilic CDs could be used in biphasic aqueous organometallic processes that are less demanding in their selectivity, such as hydrogenation reactions or carbon–carbon coupling.

Author Contributions

H.B., E.M., and S.T. conceived and designed the experiments; A.C. and H.B. performed the experiments; H.B., E.M., and S.T. analyzed the data; F.D.-P. contributed reagents/materials and analysis; E.M. and S.T. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was performed in partnership with SAS PIVERT within the framework of the French Institute for Energy Transition (Institut pour la Transition Energétique (ITE) P.I.V.E.R.T.; http://www.institut-pivert.com as part of Investment for the Future (Investissements d’Avenir). This work was supported by the French Government as part of Investments for the Future under reference ANR-001-01.

Acknowledgments

Jérémy Ternel, Nicolas Kania, Dominique Prévost, and Johan Hachani are acknowledged for the NMR and MALDI-MS analyses, and for the technical assistance. PIVERT is acknowledged for the financial support of A.C.’s thesis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Szejtli, J. Introduction and general overview of cyclodextrin chemistry. Chem. Rev. 1998, 98, 1743–1753. [Google Scholar] [CrossRef] [PubMed]
  2. Crini, G. Review: A history of cyclodextrins. Chem. Rev. 2014, 114, 10940–10975. [Google Scholar] [CrossRef]
  3. Hedges, A.R. Industrial applications of cyclodextrins. Chem. Rev. 1998, 98, 2035–2044. [Google Scholar] [CrossRef] [PubMed]
  4. Tilloy, S.; Binkowski-Machut, C.; Menuel, S.; Bricout, H.; Monflier, E. Phosphane-Based Cyclodextrins as Mass Transfer Agents and Ligands for Aqueous Organometallic Catalysis. Molecules 2012, 17, 13062–13072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Hapiot, F.; Bricout, H.; Tilloy, S.; Monflier, E. Functionalized Cyclodextrins as First and Second Coordination Sphere Ligands for Aqueous Organometallic Catalysis. Eur. J. Org. Chem. 2012, 2012, 1571–1578. [Google Scholar] [CrossRef]
  6. Tilloy, S.; Bricout, H.; Menuel, S.; Hapiot, F.; Monflier, E. Cyclodextrins Modified by Metal-Coordinating Groups for Aqueous Organometallic Catalysis: What Remains to be Done? Curr. Organocatal. 2016, 3, 24–31. [Google Scholar] [CrossRef]
  7. Reetz, M.T.; Waldvogel, S.R. β-Cyclodextrin-Modified Diphosphanes as Ligands for Supramolecular Rhodium Catalysts. Angew. Chem. Int. Ed. 1997, 36, 865–867. [Google Scholar] [CrossRef]
  8. Armspach, D.; Matt, D. Metal-capped α-cyclodextrins: The crowning of the oligosaccharide torus with precious metals. Chem. Commun. 1999, 12, 1073–1074. [Google Scholar] [CrossRef]
  9. Guitet, M.; Marcelo, F.; de Beaumais, S.A.; Zhang, Y.M.; Jimenez-Barbero, J.; Tilloy, S.; Monflier, E.; Menand, M.; Sollogoub, M. Diametrically Opposed Carbenes on an α-Cyclodextrin: Synthesis, Characterization of Organometallic Complexes and Suzuki-Miyaura Coupling in Ethanol and in Water. Eur. J. Org. Chem. 2013, 18, 3691–3699. [Google Scholar] [CrossRef]
  10. Dindulkar, S.D.; Jeong, D.; Kim, H.; Jung, S. Functionalized beta-cyclodextrin as supramolecular ligand and their Pd(OAc)(2) complex: Highly efficient and reusable catalyst for Mizoroki-Heck cross-coupling reactions in aqueous medium. Carbohydr. Res. 2016, 430, 85–94. [Google Scholar] [CrossRef]
  11. Khan, R.I.; Pitchumani, K. A pyridinium modified beta-cyclodextrin: An ionic supramolecular ligand for palladium acetate in C-C coupling reactions in water. Green Chem. 2016, 18, 5518–5528. [Google Scholar] [CrossRef]
  12. Khan, R.I.; Pitchumani, K. Water-Soluble Palladium Complex of N’-(pyridin-2-yl)propane-1,3-diamine modified beta-Cyclodextrin: An efficient Catalyst for Transfer Hydrogenation of Carbonyl Compounds. ACS Sustain. Chem. Eng. 2018, 6, 16130–16138. [Google Scholar] [CrossRef]
  13. Sak, H.; Mawick, M.; Krause, N. Sustainable Gold Catalysis in Water Using Cyclodextrin-tagged NHC-Gold Complexes. ChemCatChem. 2019, 11, 5821–5829. [Google Scholar] [CrossRef] [Green Version]
  14. Bricout, H.; Hapiot, F.; Ponchel, A.; Tilloy, S.; Monflier, E. Chemically Modified Cyclodextrins: An Attractive Class of Supramolecular Hosts for the Development of Aqueous Biphasic Catalytic Processes. Sustainability 2009, 1, 924–945. [Google Scholar] [CrossRef] [Green Version]
  15. Hapiot, F.; Menuel, S.; Bricout, H.; Tilloy, S.; Monflier, E. Recent developments in cyclodextrin-mediated aqueous biphasic hydroformylation and Tsuji–Trost reactions. Appl. Organomet. Chem. 2015, 29, 580–587. [Google Scholar] [CrossRef]
  16. Hapiot, F.; Leclercq, L.; Azaroual, N.; Fourmentin, S.; Tilloy, S.; Monflier, E. Rhodium-catalyzed hydroformylation promoted by modified cyclodextrins: Current scope and future developments. Curr. Org. Synth. 2008, 5, 162–172. [Google Scholar] [CrossRef]
  17. Matsinha, L.C.; Siangwata, S.; Smith, G.S.; Makhubela, B.C.E. Aqueous biphasic hydroformylation of olefins: From classical phosphine-containing systems to emerging strategies based on water-soluble nonphosphine ligands. Catal. Rev. 2019, 61, 111–133. [Google Scholar] [CrossRef]
  18. Hapiot, F.; Bricout, H.; Tilloy, S.; Monflier, E. Hydroformylation in Aqueous Biphasic Media Assisted by Molecular Receptors. In Hydroformylation for Organic Synthesis; Taddei, M., Mann, A., Eds.; Topics in Current Chemistry; Springer: Berlin/Heidelberg, Germany, 2013; Volume 342, pp. 49–78. [Google Scholar]
  19. Hapiot, F.; Leclercq, L.; Azaroual, N.; Fourmentin, S.; Tilloy, S.; Monflier, E. Rhodium-Catalyzed Hydroformylation Promoted by Modified Cyclodextrins: Current Scope and Future Developments. Adv. Org. Synth. 2013, 14, 36–63. [Google Scholar]
  20. Bricout, H.; Hapiot, F.; Ponchel, A.; Tilloy, S.; Monflier, E. Cyclodextrins as Mass Transfer Additives in Aqueous Organometallic Catalysis. Curr. Org. Chem. 2010, 14, 1296–1307. [Google Scholar] [CrossRef]
  21. Monflier, E.; Fremy, G.; Castanet, Y.; Mortreux, A. Molecular Recognition between Chemically Modified β-Cyclodextrin and Dec-1-ene: New Prospects for Biphasic Hydroformylation of Water-Insoluble Olefins. Angew. Chem. Int. Ed. 1995, 34, 2269–2271. [Google Scholar] [CrossRef]
  22. Sieffert, N.; Wipff, G. Importance of Interfacial Adsorption in the Biphasic Hydroformylation of Higher Olefins Promoted by Cyclodextrins: A Molecular Dynamics Study at the Decene/Water Interface. Chem. Eur. J. 2007, 13, 1978–1990. [Google Scholar] [CrossRef] [PubMed]
  23. Sieffert, N.; Wipff, G. Adsorption at the Liquid−Liquid Interface in the Biphasic Rhodium Catalyzed Hydroformylation of Olefins Promoted by Cyclodextrins: A Molecular Dynamics Study. J. Phys. Chem. B 2006, 110, 4125–4134. [Google Scholar] [CrossRef] [PubMed]
  24. Leclercq, L.; Bricout, H.; Tilloy, S.; Monflier, E. Biphasic aqueous organometallic catalysis promoted by cyclodextrins: Can surface tension measurements explain the efficiency of chemically modified cyclodextrins? J. Colloid Interface Sci. 2007, 307, 481–487. [Google Scholar] [CrossRef] [PubMed]
  25. Monflier, E.; Tilloy, S.; Fremy, G.; Castanet, Y.; Mortreux, A. A further breakthrough in biphasic, rhodium-catalyzed hydroformylation: The use of per(2,6-di-O-methyl)-beta-cyclodextrin as inverse phase transfer catalyst. Tetrahedron Lett. 1995, 36, 9481–9484. [Google Scholar] [CrossRef]
  26. Fu, H.; Li, M.; Chen, H.; Li, X. Higher olefin hydroformylation in organic/aqueous biphasic system accelerated by double long-chain cationic surfactants. J. Mol. Catal. A Chem. 2006, 259, 156–160. [Google Scholar] [CrossRef]
  27. Desset, S.L.; Reader, S.W.; Cole-Hamilton, D.J. Aqueous-biphasic hydroformylation of alkenes promoted by “weak” surfactants. Green Chem. 2009, 11, 630–637. [Google Scholar] [CrossRef]
  28. Pogrzeba, T.; Müller, D.; Hamerla, T.; Esche, E.; Paul, N.; Wozny, G.; Schomäcker, R. Rhodium-Catalyzed Hydroformylation of Long-Chain Olefins in Aqueous Multiphase Systems in a Continuously Operated Miniplant. Ind. Eng. Chem. Res. 2015, 54, 11953–11960. [Google Scholar] [CrossRef]
  29. Illner, M.; Müller, D.; Esche, E.; Pogrzeba, T.; Schmidt, M.; Schomäcker, R.; Wozny, G.; Repke, J.-U. Hydroformylation in Microemulsions: Proof of Concept in a Miniplant. Ind. Eng. Chem. Res. 2016, 55, 8616–8626. [Google Scholar] [CrossRef]
  30. Bibouche, B.; Peral, D.; Stehl, D.; Söderholm, V.; Schomäcker, R.; von Klitzing, R.; Vogt, D. Multiphasic aqueous hydroformylation of 1-alkenes with micelle-like polymer particles as phase transfer agents. RSC Adv. 2018, 8, 23332–23338. [Google Scholar] [CrossRef] [Green Version]
  31. Cocq, A.; Rousseau, C.; Bricout, H.; Oliva, E.; Bonnet, V.; Djedaini-Pilard, F.; Monflier, E.; Tilloy, S. Oleic Acid Based Cyclodextrins for the Development of New Hydrosoluble Amphiphilic Compounds. Eur. J. Org. Chem. 2019, 1236–1241. [Google Scholar] [CrossRef]
  32. Cocq, A.; Rousseau, C.; Bricout, H.; Oliva, E.; Bonnet, V.; Djedaïni-Pilard, F.; Monflier, E.; Tilloy, S. Highly Water-Soluble Amphiphilic Cyclodextrins Bearing Branched and Cyclic Oleic Grafts. Eur. J. Org. Chem. 2019, 4863–4868. [Google Scholar] [CrossRef]
  33. Mathivet, T.; Meliet, C.; Castanet, Y.; Mortreux, A.; Caron, L.; Tilloy, S.; Monflier, E. Rhodium catalyzed hydroformylation of water insoluble olefins in the presence of chemically modified beta-cyclodextrins: Evidence for ligand-cyclodextrin interactions and effect of various parameters on the activity and the aldehydes selectivity. J. Mol. Catal. A Chem. 2001, 176, 105–116. [Google Scholar] [CrossRef]
  34. Gärtner, B.; Cornils, H.; Springer, P. Lappe. DE Patent 3235030, 22 September 1982. [Google Scholar]
Scheme 1. Structure and synthesis of oleic succinyl-cyclodextrins (OS-CDs).
Scheme 1. Structure and synthesis of oleic succinyl-cyclodextrins (OS-CDs).
Catalysts 10 00056 sch001
Figure 1. Matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) spectrum of β-CD-(OSG–Me)1.6 after treatment with Rh(acac)(CO)2/TPPTS catalytic system (1 and 5 molar equivalents with respect to CD). Experiment conditions: pressure, 50 bar of CO/H2 (1:1); solution, 6 mL H2O for every 21 µmol of cyclodextrin (CD); temperature, 80 °C; time, 6 h.
Figure 1. Matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) spectrum of β-CD-(OSG–Me)1.6 after treatment with Rh(acac)(CO)2/TPPTS catalytic system (1 and 5 molar equivalents with respect to CD). Experiment conditions: pressure, 50 bar of CO/H2 (1:1); solution, 6 mL H2O for every 21 µmol of cyclodextrin (CD); temperature, 80 °C; time, 6 h.
Catalysts 10 00056 g001
Table 1. Physicochemical properties of OS-CDs.
Table 1. Physicochemical properties of OS-CDs.
EntryOS-CDSolubility at 20 °C (mM)CAC at 20 °C (µM)Surface Tension 1 (mN/m)
1α-CD-(OSG–Me)1.6521533
2α-CD-(OSG–Na)1.612120636
3β-CD-(OSG–Me)1.6281139
4β-CD-(OSG–Na)1.61589438
5γ-CD-(OSG–Me)1.843830
6γ-CD-(OSG–Na)1.417619039
8RAME-β-CD-(OSG–Me)0.92472432
9RAME-β-CD-(OSG–Na)1.02881835
10HP-β-CD-(OSG–Me)1.71464636
11HP-β-CD-(OSG–Na)1.61785452
1 Surface tension value at the CAC concentration and at 20 °C.
Table 2. Rhodium-catalyzed hydroformylation of 1-decene and 1-hexadecene in presence of various promotors.1
Table 2. Rhodium-catalyzed hydroformylation of 1-decene and 1-hexadecene in presence of various promotors.1
Catalysts 10 00056 i001
EntrySubstratePromotorPromotor (/Rh)Time (h)Conversion 2 (%)Selectivity 3 (%)l/b
1C10H20(-)(-)63652.7
2C10H20β-CD-(OSG–Me)1.614100612.3
3C10H20β-CD-(OSG–Me)1.61293592.4
4C10H20β-CD-(OSG–Me)1.610.585582.6
5C10H20β-CD166802.6
6C10H20RAME-β-CD1621902.1
7C10H20(4)1.6620532.7
8C10H20β-CD + (4)1 + 1.6645532.5
9C16H32(-)(-)61622.7
10C16H32β-CD-(OSG–Me)1.61665522.8
11C16H32β-CD163782.4
12C16H32RAME-β-CD169902.1
13C16H32(4)1.6610532.7
14C16H32β-CD + (4)1 + 1.6623532.5
15C16H32β-CD-(H2–OSG–Me)1.61668582.6
16C16H32β-CD-(OSG–Me)1.6 41663502.8
17C16H32β-CD-(OSG–Me)1.6 51667512.8
1 Experiment conditions: Rh(acac)(CO)2, 21 μmol (1 eq.); trisulfonated triphenylphosphine (TPPTS), 5 eq.; substrate, 500 eq.; water, 6 mL; temperature, 80 °C; pressure, 50 bar CO/H2 (1:1); stirring rate, 1500 rpm. 2 Olefin conversion. 3 Aldehyde selectivity—(moles of aldehydes)/(moles of converted olefins) × 100. 4 Experiment conducted using aqueous phase of Experiment 10 (first recycling). 5 Experiment conducted using the aqueous phase of Experiment 16 (second recycling).
Table 3. Rhodium-catalyzed hydroformylation of 1-hexadecene in presence of various CDs. 1
Table 3. Rhodium-catalyzed hydroformylation of 1-hexadecene in presence of various CDs. 1
Catalysts 10 00056 i002
EntryPromotorEq/RhConversion 2 (%)Selectivity 3 (%)l/b
1α-CD-(OSG–Me)1.6157612.5
2β-CD-(OSG–Me)1.6165522.8
3γ-CD-(OSG–Me)1.6139513.0
4RAME-β-CD-(OSG–Me)0.9194532.5
5HP-β-CD-(OSG–Me)1.7166552.5
6α-CD-(OSG–Na)1.6115622.9
7β-CD-(OSG–Na)1.6122581.8
8γ-CD-(OSG–Na)1.419652.9
9RAME-β-CD-(OSG–Na)1.0131622.6
10HP-β-CD-(OSG–Na)1.6132662.6
11RAME-β-CD19902.1
12RAME-β-CD1018902.2
1 Experiment conditions: Rh(acac)(CO)2, 21 μmol (1 eq.); TPPTS, 5 eq.; substrate, 500 eq.; water, 6 mL; temperature, 80 °C; pressure, 50 bar CO/H2 (1:1); stirring rate, 1500 rpm; reaction time, 6 h. 2 1-hexadecene conversion. 3 Aldehyde selectivity—(moles of aldehydes)/(moles of converted olefins) × 100.

Share and Cite

MDPI and ACS Style

Cocq, A.; Bricout, H.; Djedaïni-Pilard, F.; Tilloy, S.; Monflier, E. Rhodium-Catalyzed Aqueous Biphasic Olefin Hydroformylation Promoted by Amphiphilic Cyclodextrins. Catalysts 2020, 10, 56. https://doi.org/10.3390/catal10010056

AMA Style

Cocq A, Bricout H, Djedaïni-Pilard F, Tilloy S, Monflier E. Rhodium-Catalyzed Aqueous Biphasic Olefin Hydroformylation Promoted by Amphiphilic Cyclodextrins. Catalysts. 2020; 10(1):56. https://doi.org/10.3390/catal10010056

Chicago/Turabian Style

Cocq, Aurélien, Hervé Bricout, Florence Djedaïni-Pilard, Sébastien Tilloy, and Eric Monflier. 2020. "Rhodium-Catalyzed Aqueous Biphasic Olefin Hydroformylation Promoted by Amphiphilic Cyclodextrins" Catalysts 10, no. 1: 56. https://doi.org/10.3390/catal10010056

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