Next Article in Journal
Unique Features of a New Baeyer–Villiger Monooxygenase from a Halophilic Archaeon
Previous Article in Journal
Recent Advances in Niobium-Based Materials for Photocatalytic Solar Fuel Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Additive-Free Baeyer–Villiger Oxidation of Cyclic Ketone Catalyzed by Carboxylic-Functionalized Poly(Ionic Liquids) and Polyoxometalate Ionic Self-Assemblies

1
Department of Materials Engineering, Ocean College, Minjiang University, Fuzhou 350108, China
2
Fujian Engineering and Research Centre of New Chinese Lacquer Material, Ocean College, Minjiang University, Fuzhou 350108, China
3
Fujian Provincal University Engineering Research Centre of Green Materials and Chemical Engineering, Ocean College, Minjiang University, Fuzhou 350108, China
4
Faculty of Science, Engineering and Technology, Swinburne University of Technology, Hawthorn, VIC 3122, Australia
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(1), 127; https://doi.org/10.3390/catal10010127
Submission received: 10 December 2019 / Revised: 10 January 2020 / Accepted: 11 January 2020 / Published: 16 January 2020

Abstract

:
Two carboxylic functionalized poly(ionic liquids) and polyoxometalate ionic self-assemblies with tunable amphiphilic and oxidative activity and porous structure were designed and synthesized. They were used as efficient heterogeneous catalysts exhibiting excellent catalytic activity and quite stable reusability in additive–free Baeyer-Villiger oxidation of cyclic ketones, directly using 30% H2O2 solution in liquids–phase. Under optimum conditions, the corresponding lactones were obtained with yields ranging from 59.6 to 90.5%. The unique combination of carboxylic group with polyoxometalate anion, amphiphilic property, and porous structure is revealed to be responsible for their excellent catalytic performances in oxidation.

Graphical Abstract

1. Introduction

Recently, the growing concerns about plastic’s environmental footprint and its manufacture from bio-renewable feedstocks have received considerable attention from industry and academic communities [1,2]. One of the most rapidly developing fields is aliphatic polyester with excellent hydrolytic degradability [3,4] like polylactones and polylactides, which have been widely applied in biomedicine and engineering fields. Since the efficient synthetic pathway for the preparation of aliphatic polyester is ring-opening polymerization of lactone monomer, Baeyer–Villiger (BV) oxidation represents a powerful synthetic tool for the construction of lactone from the corresponding ketone. It is carried out in no catalytic ways using organic peracids, such as m-chloroperbenzoic [5] and perdecanoic acid [6] in more than stoichiometric amounts in complex solutions, which are shock sensitive, explosive, with inadvertently low reactivity and selectivity in oxidation, and which generate equimolar carboxylic acid as waste. Therefore, there is a compelling demand for the development of more sustainable and practical BV methods using molecular oxygen or H2O2 as green oxidant. Due to the lower oxidation potential of H2O2, the catalyst is usually required to promote oxidation via carbonyl or H2O2 activation. For example, heterogeneous catalysts based on Lewis acidic transition metal (like Sn [7,8], Fe [9], Mg [10], Mn [11]), sold acid [12], sulfonic acid functionalized silica nanoparticle [13], carboxylic multi-walled carbon nanotube [14], and macroporous CNT-sieve-composite-supported polyoxometalate [15] have been reported; however, these methods are confronted with the drawbacks of low activity or selectivity and complicated preparation processes, using high concentrations of H2O2 and so on.
Polymeric ionic liquids (PILs), as a kind of designable multifunctional polyelectrolyte, combine the properties of polymer and ionic liquid. By changing the anion and cation, a variety of functional materials with unique physicochemical properties and catalytic activities have been fabricated [16]. For example, introducing various polyoxometalate (POM) anions to provide a series of PILs-based POM composites or hybrids, these compounds acted as green heterogeneous catalysts with high catalytic activity and selectivity in olefins epoxidation with H2O2 [17], electrocatalytic water oxidation [18], and sulfides selective oxidation [19], starting from SO3H functionalized acidic ionic liquids monomer via controllable polymerization, to prepare microporous, mesoporous, microporous, and hierarchically nanoporous PILs with strong acid strength and use them as efficient solid acids in various acid-catalyzed reactions [20,21,22,23]. Nevertheless, the self-assembled POM and PILs composites or hybrids are always nonporous. The hydrophilic or hydrophobic frameworks of polymeric cation usually hinder the interaction of the POM catalytic center with the hydrophobic substrate and limit the heterogeneous catalytic activity in mass-transfer control.
Aiming to develop a green and sustainable protocol for BV oxidation with H2O2, in this context, two novel multifunctional PILs and POM ionic self-assemblies (denoted as COOH-PIL@P-POM and COOH-PIL@Si-POM) (Scheme 1) with tunable amphiphilic and oxidative properties were fabricated by ionic self-assembling of carboxylic functionalized PILs with Keggin type H3PW12O40 and H3PW12O40 in water. The as-synthesized ionic self-assemblies served as efficient heterogeneous catalysts exhibiting excellent catalytic activity and stable reusability in additive–free BV oxidation of cyclic ketones in the presence of 30% H2O2 under liquids–phase. The corresponding lactones were obtained with yields ranging from 59.6% to 90.5%. Two ionic self-assemblies can be reused for five times after simple work-up and without noticeable decrease of activity.

2. Results and Discussion

2.1. Synthesis and Characterization of Two Ionic Self-Assemblies

COOH-PIL@P-POM and COOH-PIL@Si-POM were synthesized using 2-(dimethylamino)ethyl methyacrylate, chloroacetic acid, H3PW12O40 and H4SiW12O40 as starting materials via quaternization, free radical polymerization, and ionic self-assembly in water. 2-(dimethylamino)ethyl methyacrylate as a kind of industrial available chemical shows strong basicity;this nature causes the quaternization to be carried out easily and to proceed completely under mild conditions. Importantly, the resulting ionic liquid monomer not only can be easily polymerized but also produce a polymer with linear long chain quaternary ammonium structure. The reason for choosing H3PW12O40, H4SiW12O40 as the suitable counter POM lies in: (1) two POM exhibited good performance in catalytic activity in BV oxidation via generate peroxo intermediate in situ with H2O2, and their oxidative activity could be regulated by the acidic content and acidity [15] and (2) two POM anions have the size of 1nm × 1 nm, together with their high negative charge density, which is beneficial for the adequate coating of PILs on their surfaces and the formation of an excellent amphiphilic microenvironment. On the other hand, two POM anions could function as pillar type nodes or crosslinking agents embedded and supported in the 2D net-like structure formed by the linear long chain polymeric quaternary ammonium cation and thereby create an open 3D network structure with hierarchical porosity. In addition, the synthesized carboxylic functionalized PILs and two POMs both have excellent solubility in water, but the resulting ionic self-assemblies are insoluble. Therefore, the self-assembly in water ensures a product with high purity. The produced two carboxylic functionalized ionic self-assemblies show oxidative, acidic, and excellent amphiphilic activity at the same time, and these properties can be well tuned by the polymerization degree and the number and type of POM anion.
Figure 1a shows the XRD patterns of two ionic self-assemblies; these samples exhibit two broad peaks centered at 5° to 10° and 20° to 40°, which could be attributed to the characteristic peaks of Keggin type POM, suggesting the crystal structure and Keggin configuration of POM anion is still well retained. As for the facts of the noticeable decrease in the intensities and no peaks appearing in the high angel areas, they could be ascribed to the POM anions being well dispersed in the polycationic frameworks via rearrangement during the self–assembly.
Figure 1b shows the infrared data of two ionic self-assemblies and carboxylic functionalized PIL. The absorption bonds appeared at 1725, 1630, 1470, and 1140 cm−1, which could be assigned to the vibrations of C=O groups of ester and carboxylic moiety, C-N stretching, and C-O vibration, respectively, suggesting the polycationic frameworks was introduced successfully. In the wavenumber regions of 750–1200 cm−1, PW12O40−3 gave four characteristic bands at 1080, 978, 896, and 816 cm−1. SiW12O40−4 gave five characteristic bands at 1015, 972, 920, 882, and 796 cm−1. These facts combined with XRD results prove that the basic structure and geometry of the Keggin type POM anion are well-preserved. The vibration bands of POM anion in two ionic self–assemblies are slightly shifted in comparison with free POM, which result from the strong ionic interaction between the polymeric quaternary ammonium cation and POM anion. In addition, compared with the fresh ionic self-assemblies, there is no notable difference shown in the FT-IR spectrum of the used ionic self-assemblies after five recycles, indicating the highly chemical and structural stability of two ionic self-assemblies.
TG curves of COOH-PIL@P-POM and COOH-PIL@Si-POM are presented in Figure 2. All samples show similar changing trend from 30 °C to 400 °C. There is an initial weight loss of about 1% to 1.5% that can be observed up to 100 °C, which is attributed to the adsorbed moisture and hydrated water of POM anions and is almost stable up to 250 °C (the loss of weigh less than 5%). The main weight loss occurred around 250 °C, which indicates the slow decomposition of the counter polymeric organic cation.
SEM images of two self-assemblies with different viscosity average molecular weight were shown in Figure 3a–d. As it can be seen, the prepared samples displayed the aggregation morphology of small spherical and block nanoparticles with a size of about hundreds of nanometers. These nanoparticles further packed and linked each other to form irregular secondary particles with the size of a micrometer. It is obvious that abundant net-like hierarchical mesopores and macropores can be found in the secondary particles. All samples have waxy smooth surface without observable pores, indicating the coating of polymeric quaternary ammonium cation around the surface of POM anion is sufficient and uniform, and POM anions are completely entrapped. The effect of polymeric quaternary ammonium cation on the morphology of ionic self-assembly is remarkable. With an increase in polymerization degree, the size of particle improved, due to the agglomeration of polymeric cation on the surface of POMs anion. When the POM anion changed from PW12O40−3 to SiW12O40−4, there was a remarkable improvement in degree of cross-linking between particles, which could be ascribed to the increase of negative charge of POM anion. In addition, no separate aggregation of PILs was found, indicating that the PILs coating has occurred only on the surface of POM anion.
Figure 4 shows the N2 adsorption–desorption isotherms and the corresponding pore size distribution curves of COOH-PIL@P-POM. According to Brunauer–Deming–Deming–Teller classification, these isotherms can be categorized as type IV, indicating the presence of mesopores in ionic self-assembly. Pore diameter curve manifests wide pore size distribution from 5 nm to 120 nm and a broad peak centered at ca. 30–60 nm. These pores located at the empty space inside the hierarchical architectures, which is consistent with the results of SEM. The BET specific surface area of COOH-PIL@P-POM is 26.46 m3/g. The high specific surface area and hierarchical pore structure are favorable for COOH-PIL@P-POM used as effective heterogeneous catalysis in BV oxidation.

2.2. Catalytic Performance in Additive-Free BV Oxidation

Since ε-caprolactone is the most important lactone used in industry with global consumption of over 30 k ton annually, initially, the oxidation of cyclohexanone under atmospheric pressure catalyzed by two ionic self-assemblies was evaluated as model reaction. Various reaction parameters including the amount of catalyst, temperature, the molar ratio of H2O2 to ketone, time, and solvent were investigated. Under optimum conditions, these results were listed in Table 1. As shown in Table 1 (entries 5–8), both ionic self-assemblies exhibited notable advantages of lower usage of catalysts, high conversion and selectivity, shorter reaction time, and lower reaction temperature.
These facts demonstrate the existence of synergistic catalytic effect between the POM anion and carboxylic functionalized polymeric cation. During the oxidation, the abundant carboxylic acid group in polymeric cation framework and POM anion could be transferred into active peroxide acid and POM peroxo complex in situ with H2O2, respectively. They acted as active oxidants oxidize the cyclic ketone into product lactone. Moreover, these better results also associated with the mesopore structure with acidic and amphiphilic activity surface, which distributed on the surface of two ionic self-assemblies. On the one hand, they facilitated the adsorption and transfer of reactants and supplied a microenvironment for the evolution of ketone and H2O2 molecules into the catalyst pore channels. On the other hand, the formed product lactone could be separated immediately, avoiding the over-oxidation and hydrolysis under acidic surroundings. Under the same COOH-PIL conditions, COOH-PIL@P-POM gave better results than COOH-PIL@Si-POM. The reasons lie in the following facts: (1) The higher negative charge on the surface of SiW12O40−4 caused the obtained ionic self-assembly processes’ higher acidity content, which increased the probability of hydrolysis of lactone. Meanwhile, this feature could enhance the cross-linking between nanoparticles, and as a consequence, the number of pores on the surface of ionic self-assembly decreased, and the pore cavity became smaller. (2) PW12O40−3 based peroxo POM anion species presented higher oxidative activity than that of SiW12O40−4. The influence of polymeric organic anion on BV oxidation was also remarkable. With the increasing of FW, the conversion of oxidation significantly improved, since the number of carboxylic acid groups and the count of POM anions increased correspondingly. There was a slight decline in selectivity due to the enhancement of oxidative catalytic activity. Moreover, the crosslinking surface morphology of ionic self-assemblies resulting from the increasing of FW could impede oxidation proceeding via the limitation of mass transfer of substrate to the active site on the micropore channel. In order to obtain the best oxidation results, an optimization between the synergistic catalytic effect and morphology control should be achieved. The scope and limitation of cyclic ketone was also investigated, and the results were summarized in Table 2. All the selected substrates reacted smoothly and gave the corresponding lactones with good to excellent yields. Among them, 2-adamantanone gave the best result for its large molecular size and rigid structure, which makes it difficult to hydrolysis and leads to high chemoselectivity of the corresponding lactone. Owing to the higher ring strain and lower steric hindrance, cyclobutanone and cyclopentanone gave higher yields than cyclohexanone. Additionally, no substituent effect was observed for mono substituted cyclohexanone, with methyl at the para-, meta-, and ortho-position found to have a minimal effect on the outcome of the oxidation. A little surprise in this work is the oxidation of cycloheptanone; the corresponding lactone was obtained with a yield of 7%. This value is almost double compared with the literature [6]. To demonstrate the scalability and synthetic utility of this present protocol, a hundred-gram scale oxidation of cyclohexanone (1 mol) was carried out. Under the standard conditions, ε-caprolactone was obtained in the yield of 74%.

2.3. Catalyst Reusability

The reusability of two self-assemblies was then investigated, using cyclopentanone as the model substrate. After the first use, two ionic self-assemblies were separated directly by simple filtration and washed with ethyl ester, dried under vacuum, and subjected to another cycle with a fresh reaction mixture. The result was shown in Figure 5. Two self-assemblies could be quantitatively recovered from the reaction mixture and reused consecutively for five times without significant loss in catalytic performance compared with that of the fresh catalyst. Moreover, the TG curve, XRD, IR spectra, and elemental analysis of recycled composites were also determined and found to be very close to the data for fresh catalyst. In addition, ICP analysis of the solution after reaction showed that the POM anion contents were negligible, also indicating that no leaching of active component occurred during the reaction. The high stability and recyclability of this catalytic system offers a potential candidate for an economic and environmentally friendly process for synthesis of lactones in future.

3. Materials and Methods

3.1. Chemicals and Materials

All chemicals and solvents were commercially available and used as received without further purification. FT-IR spectra were recorded with a Bruker Tensor-27 spectrometer (Ettlinggen, BW, Germany) in the range 400–4000 cm−1. Powder X-ray diffraction (XRD) were studied by a with Cu anode for scanning ranged from 0° to 90° of 2θ using a Rigaku D/Max 2500 diffractometer (Tokyo, Japan). Scanning electron microscopy (SEM) was studied with a Hitachi S-4800 microscope (Tokyo, Janpan). C, H, and N elemental analyses were performed on a Vario EL III elemental analyzer (Hanau, HD, Germany). The thermal analysis (TGA) was performed using a TGA Q 50 instrument (Newcastle, DE, USA); samples were heated at a constant heating rate of 10 °C.min−1 room temperature to 400 °C under N2 atmosphere. Nuclear magnetic resonance (NMR) spectra were recorded on a Bruker AC-400 spectrometer (Ettlinggen, BW, Germany) in appropriate deuterated solvents.

3.2. Preparation of Ionic Liquid Monomer (COOH-PIL)

To a mixture of 2-(dimethylamino) ethyl methyacrylate (4.8 g, 30 mmol) in 30 mL acetone, 1-chloro acetic acid (3.0 g, 30 mmol) was added in dropwise, and it was stirred at 65 °C for 24 h under the protection of nitrogen. On completion, solvent was removed by rotary evaporation. After drying under vacuum, COOH-IL was obtained with a yield of 95%. FT-IR(KBr, cm−1): 3567, 3010, 2959, 1728, 1631, 1473, 1390, 1337, 1147. 1 H NMR (400 MHz, D2O): δ = 6.09 (s, 1H), 5.68 (s, 1H), 4.51 (t, J = 6.8 Hz, 2H, O-CH2), 4.00 (m, J = 1.5 Hz, 2H, -CH2), 3.65 (m, 2H, -CH2), 3.21(t, 6H, 2-CH3), 1.98 (s, 3H, -CH3) (Figure S1). Anal. Calcd. for C10H18N1O4Cl: C, 47.67; H, 7.15; N, 5.15. Found: C, 47.70; H, 7.24; N, 5.11.
Polymerization was carried out with AIBA (2,2′-azobis(2-methylpropionamide) dihydrochloride) as the initiator in a degassed tube under N2. In a typical procedure, ionic liquid monomer COOH-IL (3 g, 16 mmol), 15 mL deionized water, and 7 mL EDTA aqueous solution were charged; the mixture was degassed by bubbling N2 to remove oxygen and was heated to 50 °C. After addition of 15 mL AIBA aqueous solution (0.27 g, 1.25 mmol) in dropwise, the polymerization was carried out at 65 °C for 12 h under N2 atmosphere. On completion, the polymers were purified by sedimentation from cold acetone and dried under vacuum at 80 °C overnight. A total of 2.8 g COOH-functionalized PILs (COOH-PIL) was obtained in the yield of 91%. FT-IR(KBr, cm−1): 3421, 2962, 1729, 1621, 1474, 1399, 1349, 1224, 1147.
1 H NMR (400MHz, D2O): δ = 4.02 (s, 2H, -CH2), 3.62 (t, 2H, -CH2), 2.65 (t, 2H, -CH2), 2.35 (s, 6H, 2CH3), 1.78 (t, 2H, CH2), 0.85 (s, 3H, -CH3) (Figure S1) Anal. Calcd. for C10nH18nNnO4nCln: C, 47.67; H, 7.15; N, 5.15. Found: C, 47.71; H, 7.19; N, 5.09. The viscosity-average molecular weight was determined by Vickers viscometer (Hangzhou, China), and the values were 15,000, 23,000, and 30,000, respectively.

3.3. Preparation of Ionic Self-Assemblies

In a typical experiment, COOH-PIL (0.5 g) was thoroughly dissolved in 20 mL deionized water; then, aqueous solution of H3PW12O40 (1.9 g, 0.45 mmol) or H4SiW12O40 (1.4 g, 0.42 mmol) was added in dropwise under vigorous stirring conditions. A precipitate was formed immediately. The suspension was stirred at room temperature for 24 h. Afterwards, the resulting ionic self-assembly was collected by filtration, washed with water, and dried overnight under vacuum at 80 °C to afford the final products with isolated yields of 89% and 91%, respectively.
COOH-PL@P-POM, FT-IR (KBr, cm−1): 3408, 2964, 1728, 1620, 1473, 1222, 1145, 1080, 979, 896, 816, 594, 507. Anal. Calcd. for C10nH18nNnO17.33nP0.33nW4n: C, 10.22; H, 1.54; N, 1.19. Found: C, 10.24; H, 1.44; N, 1.15.
COOH-PL@Si-POM, FT-IR (KBr, cm−1): 3489, 2994, 1727, 1618, 1466, 1232, 1143, 1015, 972, 921, 882, 796, 528. Anal. Calcd. for C10nH18nNnO14nSi0.25nW3n: C, 14.79; H, 1.94; N, 1.50. Found: C, 14.72; H, 1.88; N, 1.48.

3.4. BV Oxidation of Cyclic Ketones

In a typical experiment, a mixture of cyclic ketone (10 mmol), dichloroethane (15 mL), COOH-PIL@POM (0.15 g) were added into a 25 mL three-necked flask. Under the protection of nitrogen, the reaction mixture was heated up to 70 °C with vigorous stirring, and aqueous H2O2 (30 wt %, 15 mmol) was added into within 20 min. The reaction was stirred for another 4 h. On completion, the catalyst was separated from the mixture by centrifugation, and the liquid was extracted with ethyl ether (3 × 10 mL). Combined the organic layer, it was dried over anhydrous sodium sulfate and collected under vacuum distillation. The product in the resulting mixture was determined and quantitatively analyzed by a gas chromatography–mass spectrometry (GC-MS; PerkinElmer, Clarus SQ-8 equipped with a carpillary column, TC-1701(GL Sciences, length 60 m, i.d. 0.25 μm, and film thickness 0.25 μm).
The recovered catalyst was washed with ethyl ether, dried under vacuum at 80 °C, and then reused in the next run; the reaction was performed as described above. The concentration of H2O2 in its aqueous solution was determined iodometrically prior to use in the oxidation reaction. Self-decomposition of H2O2 as a function of time was monitored by measuring the volume of oxygen evolved.

4. Conclusions

In conclusion, we have developed an efficient approach for cyclic ketone additive–free BV oxidation in the presence of H2O2, using two carboxylic acid functionalized PILs and POM ionic self-assemblies with tunable amphiphilic and oxidative activity and porous structure as heterogeneous catalysts. Under optimized conditions, the target lactones were obtained with isolated yields ranging from 59.6% to 90.5%. There are several noteworthy features of this approach including ionic self-assemblies could be easily prepared using industrial available chemicals, good yield, mild reaction condition, simple workups and purification, high stable reusability of self-assemblies, etc., which make it potentially applicable in industry. This present work suggests that the combination of carboxylic acid functionalized long carbon chain cations with POM anions can work together for the BV oxidation of cyclic ketone. For carboxylic acid functionalized polymeric cation organic frameworks, which not only generated active peracids in situ but also provided a nano-sized micro-environment with amphiphilic activity, which is beneficial for H2O2 and ketone approaching the catalytic active site on pore cavity, on the other hand, the produced lactones could be removed in time, avoiding the undesirable over-oxidation and hydrolysis under acidic environment. POM anion acted as a rigid pillar node embedded and supported in the 2D frameworks formed by carboxylic functionalized polymeric quaternary ammonium cations, thereby fabricating an open three-dimensional network structure with hierarchical porosity. Meanwhile, POM anion also served as precursor for active catalytic intermediate POM peroxide species.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/1/127/s1, Figure S1: 1H NMP of COOH-IL and COOH-PIL, Figure S2: GC-MS Spectra of Cyclopentanone Oxidation Product, Figure S3: GC-MS Spectra of Cyclohexanone Oxidation Product.

Author Contributions

Conceptualization, investigation, and supervision: X.Z., Q.L., and A.Y.; proposed methodology and supervision: X.L. and H.X.; validation, formal analysis, and writing: Z.X. and H.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 51703090), Research Project of Fujian Province Education Department (No. JAT160373), the Open Project of State Key Laboratory of Structural Chemistry (20170030), and the Research Project of Minjiang University (No. MYK19010).

Acknowledgments

The authors are grateful for the funding from the National Natural Science Foundation of China, Fujian Province Education Department and Minjiang University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Badu, R.P.; Oconnor, K.; Seeram, R. Current progress on bio-based polymers and their future trends. Prog. Biomate. 2013, 2, 8. [Google Scholar]
  2. Zhu, Y.; Romain, C.; Williams, C.K. Sustainable polymers from renewable resources. Nature 2016, 540, 352–362. [Google Scholar] [CrossRef]
  3. Hillmyer, M.A.; Tolman, W.B. Aliphatic polyester block polymers: Renewable, degradable, and sustainable. Acc. Chem. Res. 2014, 47, 2390–2396. [Google Scholar] [CrossRef]
  4. Hillmyer, A.C.; Tolman, W.B. Aliphatic polyesters: Synthesis, properties and applications. Adv. Polym. Sci. 2002, 157, 1–40. [Google Scholar]
  5. Tenbrink, G.J.; Arends, I.W.C.E.; Sheldon, R.A. The-Baeyer-Villiger reaction: New developments toward greener procedures. Chem. Rev. 2004, 104, 4105–4124. [Google Scholar] [CrossRef] [PubMed]
  6. Sitko, M.; Szelwicka, A.; Wojewodka, A.; Skwarek, A.; Tadasiewicz, D.; Schimmelpfennig, L.; Dziuba, K.; Witczakc, M.M.; Chrobok, A. Perdecanoic acid as a safe and stable mediumchain peracid for Baeyer–Villiger oxidation of cyclic ketones to lactones. RSC Adv. 2019, 9, 30012–30018. [Google Scholar] [CrossRef] [Green Version]
  7. Corma, A.; Navarro, M.T.; Renz, M. Lewis acidic Sn (IV) centers-grafted onto MCM-41 as catalytic sites for the Baeyer–Villiger oxidation with hydrogen peroxide. J. Catal. 2003, 219, 242–246. [Google Scholar] [CrossRef]
  8. Chen, T.; Wang, B.; Li, Y.; Liu, L.; Qiu, S. Hydrothermal synthesis of tin contain mesoporous silicas and their catalytic performance over Baeyer–Villiger oxidation of cyclohexanone to ε-carpolactone: Comparison of Sn/MCM-41 and Sn/SBA-15. J. Porrous Mater. 2015, 22, 949–957. [Google Scholar] [CrossRef]
  9. Mandai, K.; Hanata, M.; Mitsudo, K.; Mandai, H.; Hashimoto, H.; Takada, J. Bateriogenic iron oxide as an effective catalyst for Baeyer-Villiger oxidation with molecular oxygen and benzaldehyde. Tetrahedron 2015, 71, 9403–9407. [Google Scholar] [CrossRef]
  10. Llamas, R.; Jimenez-Sanchidrian, C.; Ruiz, J.R. Environmentally friendly Baeyer-Villiger oxidation with H2O2/nitrile over Mg(OH)2 and MgO. Appl. Catal. B Environ. 2007, 72, 18–25. [Google Scholar] [CrossRef]
  11. Thomas, J.M.; Raja, R. Catalytically active centres in porous oxides: Design and performance of highly selective news catalysts. Chem. Commun. 2001, 8, 675–687. [Google Scholar] [CrossRef]
  12. Fischer, J.; Holderich, W.F. Baeyer–Villiger-oxidation of cyclopentanone with aqueous hydrogen by acid heterogeneous catalysis. Appl. Catal. A Gen. 1999, 180, 435–443. [Google Scholar] [CrossRef]
  13. Xing, W.Z.; Ma, Q.G.; Peng, X.H. Silica/A153-SO3H: An efficient catalyst for the Baeyer-Villiger oxidation of cyclic ketones with hydrogen peroxide. C. R. Chim. 2015, 18, 581–585. [Google Scholar] [CrossRef]
  14. Chen, S.Y.; Zhou, X.T.; Wang, J.Z.; Luo, R.C.; Luo, Q.J.; Yu, L.J.; Ji, H.B. Promoting the aerobic Baeyer-Villiger oxidation of ketone over carboxylic multi-walled carbon namotubes. Mol. Catal. 2017, 438, 152–157. [Google Scholar] [CrossRef]
  15. Wei, Y.L.; Yang, Z.W.; Zhao, G.H.; Yang, W. Synthesis and catalytic properties of macroporpus SiO2-coated CNT-sieve-composite supported 12-tungstophosph-oric acid catalysts with dual pore structure for Baeyer-Villiger oxidation of cyclic ketones under microwave irradition. J. Catal. 2019, 371, 196–206. [Google Scholar] [CrossRef]
  16. Yuan, J.; Mecerreyes, D.; Antonietti, M. Poly(ionic liquid)s: An updaye. Prog. Polym. Sci. 2014, 38, 1009. [Google Scholar] [CrossRef]
  17. Leng, Y.; Zhao, J.; Jiang, P.; Wang, J. Amphiphilic polyoxometalate-paired polymer coated Fe3O4: Magnetically recyclable catalyst for epoxidation of bio-derived olefins with H2O2. ACS Appl. Mater. Interfaces 2014, 6, 5947–5954. [Google Scholar] [CrossRef]
  18. Adhikary, S.D.; Tiwari, A.; Nagaiah, T.C.; Manda, D. Stabilization of cobalt-poly-oxometalate over poly(ionic liquids) compsiites for efficient electrocatalytic water oxidation. ACS Appl. Mater. Interfaces 2018, 10, 38872–38879. [Google Scholar] [CrossRef]
  19. Fafiee, E.; Shahebrahimi, S. Organic–inorganic hybrid polyionic liquid based polyoxomatalate as nano porous material for selective oxidation of sulfides. J. Mole. Struct. 2017, 1139, 255–263. [Google Scholar]
  20. Kim, D.W.; Chi, D.Y. Polymer-supported ionic liquids: Imidazolium salts as catalysts for nucleophilic substitution reaction including fluorinations. Angew. Chem. Int. Ed. 2004, 43, 483–485. [Google Scholar] [CrossRef]
  21. Parvulescu, C.I.; Hardacre, C. Catalysis in ionic liquids. Chem. Rev. 2007, 107, 2615–2665. [Google Scholar] [CrossRef]
  22. Tao, D.J.; Liu, F.J.; Wang, L.; Jiang, L.L. A green and efficient hydration of alkynes catalyzed by hierarchically porous poly(ionic liquid)s. Appl. Catal. A Gen. 2018, 564, 56–63. [Google Scholar] [CrossRef]
  23. Jin, M.J.; Taher, A.; Kang, H.J.; Choi, M.; Ryoo, R. Palladium acetate immobiliz-ed in a hierarchical MFI zeolite-supported ionic liquid: A highly active and recyclable catalyst for Suzuki reaction in water. Green Chem. 2009, 11, 309–313. [Google Scholar] [CrossRef]
Scheme 1. Additive–free BV oxidation of cyclic ketones with H2O2 catalyzed by carboxylic. functionalized PILs and POM ionic self–assemblies COOH-PIL@P-POM and COOH-PIL@Si-POM.
Scheme 1. Additive–free BV oxidation of cyclic ketones with H2O2 catalyzed by carboxylic. functionalized PILs and POM ionic self–assemblies COOH-PIL@P-POM and COOH-PIL@Si-POM.
Catalysts 10 00127 sch001
Figure 1. (a) XRD patterns of COOH-PIL@P-POM and COOH-PIL@Si-POM; (b) FT-IR spectra of COOH-PIL@P-POM, COOH-PIL@Si-POM, and COOH-PIL.
Figure 1. (a) XRD patterns of COOH-PIL@P-POM and COOH-PIL@Si-POM; (b) FT-IR spectra of COOH-PIL@P-POM, COOH-PIL@Si-POM, and COOH-PIL.
Catalysts 10 00127 g001
Figure 2. TG patterns of COOH-PIL@P-POM and COOH-PIL@Si-POM. The viscosity average FW of 23,000.
Figure 2. TG patterns of COOH-PIL@P-POM and COOH-PIL@Si-POM. The viscosity average FW of 23,000.
Catalysts 10 00127 g002
Figure 3. SEM images of two ionic self-assemblies. (a) COOH-PL@P-POM, the viscosity average FW of the counter polycation is 23,000; (b) COOH-PL@Si-POM, the viscosity average FW of the counter polycation is 23,000; (c) COOH-PL@P-POM, the viscosity average FW of the counter polycation is 15,000; (d) COOH-PL@Si-POM, the viscosity average FW of the counter polycation is 15,000.
Figure 3. SEM images of two ionic self-assemblies. (a) COOH-PL@P-POM, the viscosity average FW of the counter polycation is 23,000; (b) COOH-PL@Si-POM, the viscosity average FW of the counter polycation is 23,000; (c) COOH-PL@P-POM, the viscosity average FW of the counter polycation is 15,000; (d) COOH-PL@Si-POM, the viscosity average FW of the counter polycation is 15,000.
Catalysts 10 00127 g003
Figure 4. N2 adsorption/desorption isotherms (a) and pore size distribution (b) curves of COOH-PIL@P-POM (FW = 23,000).
Figure 4. N2 adsorption/desorption isotherms (a) and pore size distribution (b) curves of COOH-PIL@P-POM (FW = 23,000).
Catalysts 10 00127 g004
Figure 5. Reusability of COOH-PIL@P-POM and COOH-PIL@Si-POM in BV oxidation of cyclopentanone.
Figure 5. Reusability of COOH-PIL@P-POM and COOH-PIL@Si-POM in BV oxidation of cyclopentanone.
Catalysts 10 00127 g005
Table 1. The catalytic performance of various catalysts in Baeyer–Villiger (BV) oxidation of cyclohexanone with H2O2.
Table 1. The catalytic performance of various catalysts in Baeyer–Villiger (BV) oxidation of cyclohexanone with H2O2.
EntryCatalystReaction ConditionsConversion (%)Selectivity (%)Yield (%)
1blank5 mmol ketone, H2O2 15 mmol, DCE 15 mL, 60 °C,4 h---
2H3PW12O405 mmol ketone, catalyst 0.0025 mmol, H2O2 15 mmol, DCE 15 mL, 60 °C, 4 h236715.4
3H4SiW12O405 mmol ketone, catalyst 0.0025 mmol, H2O2 15 mmol, DCE 15 mL, 60 °C, 4 h128410.8
4COOH@PIL5 mmol ketone, catalyst 0.0025 mmol, H2O2 15 mmol, DCE 15 mL, 60 °C, 4 h10.2828.4 a
5COOH-PIL@P-POM5 mmol ketone, catalyst 0.0025 mmol, H2O2 15 mmol, DCE 15 mL, 60 °C, 4 h83.49377.6 a
6COOH-PIL@Si-POM5 mmol ketone, catalyst 0.0025 mmol, H2O2 15 mmol, DCE 15 mL, 60 °C, 4 h68.59061.6 a
7COOH-PIL@P-POM5 mmol ketone, catalyst 0.0025 mmol, H2O2 15 mmol, DCE 15 mL, 60 °C, 4 h78.69272.3 b
8COOH-PIL@P-POM5 mmol ketone, catalyst 0.0025 mmol, H2O2 15 mmol, DCE 15 mL, 60 °C, 4 h87.38574.2 c
a COOH-PIL, FW = 23,000; b COOH-PIL, FW = 15,000; c COOH-PIL, FW = 30,000.
Table 2. Results of BV oxidation of various cyclic ketones with H2O2 catalyzed by two ionic self-assemblies.
Table 2. Results of BV oxidation of various cyclic ketones with H2O2 catalyzed by two ionic self-assemblies.
EntryCatalystKetoneLactoneYield a (%)
1COOH-PIL@ P-POM Catalysts 10 00127 i001 Catalysts 10 00127 i00286.7
2COOH-PIL@Si-POM80.3
3COOH-PIL@P-POM Catalysts 10 00127 i003 Catalysts 10 00127 i00485.2
4COOH-PIL@Si-POM78.4
5COOH-PIL@P-POM Catalysts 10 00127 i005 Catalysts 10 00127 i00690.5
6COOH-PIL@Si-POM81.3
7COOH-PIL@P-POM Catalysts 10 00127 i007 Catalysts 10 00127 i00876.2
8COOH-PIL@Si-POM60.8
9COOH-PIL@P-POM Catalysts 10 00127 i009 Catalysts 10 00127 i01075.7
10COOH-PIL@Si-POM59.6
11COOH-PIL@P-POM Catalysts 10 00127 i011 Catalysts 10 00127 i01275.8
12COOH-PIL@Si-POM60.2
13COOH-PIL@P-POM Catalysts 10 00127 i013 Catalysts 10 00127 i0147.2
14COOH-PIL@Si-POM3.6
a Reaction conditions: cyclohexanone (5 mmol); catalyst (0.0025 mmol); H2O2 (30%, 15 mmol); 15 mL DCE; temperature, 60 °C; time, 4 h.

Share and Cite

MDPI and ACS Style

Li, X.; Xue, H.; Lin, Q.; Yu, A. Additive-Free Baeyer–Villiger Oxidation of Cyclic Ketone Catalyzed by Carboxylic-Functionalized Poly(Ionic Liquids) and Polyoxometalate Ionic Self-Assemblies. Catalysts 2020, 10, 127. https://doi.org/10.3390/catal10010127

AMA Style

Li X, Xue H, Lin Q, Yu A. Additive-Free Baeyer–Villiger Oxidation of Cyclic Ketone Catalyzed by Carboxylic-Functionalized Poly(Ionic Liquids) and Polyoxometalate Ionic Self-Assemblies. Catalysts. 2020; 10(1):127. https://doi.org/10.3390/catal10010127

Chicago/Turabian Style

Li, Xinzhong, Hanyu Xue, Qi Lin, and Aimin Yu. 2020. "Additive-Free Baeyer–Villiger Oxidation of Cyclic Ketone Catalyzed by Carboxylic-Functionalized Poly(Ionic Liquids) and Polyoxometalate Ionic Self-Assemblies" Catalysts 10, no. 1: 127. https://doi.org/10.3390/catal10010127

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