Next Article in Journal
Catalytic Processes from Biomass-Derived Hexoses and Pentoses: A Recent Literature Overview
Previous Article in Journal
Characterization of Highly Dispersed Rod- and Particle-Shaped CuFe19Ox Catalysts and Their Shape Effects on WGS
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Copper(II) Complexes of Arylhydrazone of 1H-Indene-1,3(2H)-dione as Catalysts for the Oxidation of Cyclohexane in Ionic Liquids

by
Gonçalo A. O. Tiago
1,
Ana P. C. Ribeiro
1,*,
M. Fátima C. Guedes da Silva
1,
Kamran T. Mahmudov
1,2,
Luís C. Branco
3 and
Armando J. L. Pombeiro
1,*
1
Centro de Química Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisbon, Portugal
2
Department of Chemistry, Baku State University, Z. Xalilov Str. 23, Az 1148 Baku, Azerbaijan
3
REQUINTE, Departamento de Química, Faculdade de Ciências e Tecnologias da Universidade Nova de Lisboa, Quinta da Torre, 2829-516 Caparica, Portugal
*
Authors to whom correspondence should be addressed.
Catalysts 2018, 8(12), 636; https://doi.org/10.3390/catal8120636
Submission received: 6 November 2018 / Revised: 28 November 2018 / Accepted: 30 November 2018 / Published: 7 December 2018
(This article belongs to the Section Catalytic Materials)

Abstract

:
The copper(II) complexes [CuL(H2O)2]∙H2O (1) and [CuL(dea)] (2) [L = 2-(2-(1,3-dioxo-1H-inden-2(3H)-ylidene)hydrazinyl)benzenesulfonate, dea = diethanolamine] were applied as catalysts in the peroxidative (with tert-butyl-hydroperoxide or hydrogen peroxide) conversion of cyclohexane to cyclohexanol and cyclohexanone, either in acetonitrile or in any of the ionic liquids [bmim][NTf2] and [hmim][NTf2] [bmim = 1-butyl-3-methylimidazolium, hmim = 1-hexyl-3-methylimidazolium, NTf2 = bis(trifluoromethanesulfonyl) imide]. Tert-butyl-hydroperoxide led to better product yields, as compared to H2O2, with a selectivity directed towards cyclohexanone. The ILs showed a better performance than the conventional solvent for the copper complex 1. No catalytic activity was observed for 2 in the presence of an IL.

Graphical Abstract

1. Introduction

Due to their intrinsically low reactivity, the oxidation of saturated hydrocarbons under mild reaction conditions is a great challenge in catalysis [1,2,3,4,5,6,7]. This is of particular significance to the oxidation of cyclohexane to cyclohexanol and cyclohexanone, in view of the involvement of these products for the production of nylon-6 and nylon-6,6 [8,9].
The conversion of alkanes into valuable products using green oxidants such as molecular oxygen, tert-butyl hydroperoxide (TBHP) or hydrogen peroxide is attaining considerable attention, and a variety of transition metals are being applied [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21].
The usage of ionic liquids (ILs) as solvents in catalysis [22,23,24,25] and for cyclohexane oxidation in particular is scarce. Some studies with manganese(III) porphyrins as catalysts in [bmim][PF6] (PF6 = hexafluorophosphate) mixed with dichlorometane or acetonitrile showed excellent yields (up to 88%) for the hydroxylation of cyclohexane, cyclooctane, adamantane and tetralin [26], as well as of tetrahydronaphthalene [27]. Examples of transition metal catalysts for the oxidation of cyclohexane include iron scorpionates [28,29,30,31,32,33,34,35], copper(II) terpyridine-based compounds [36], benzene–sulfonate and benzene–carboxylate copper(II) polymers [37,38], and polynuclear copper(II)–arylhydrazone complexes [39], and other metals, such as vanadium [40,41,42,43,44,45,46].
Despite the potential applications of arylhydrazone and their metal complexes, mainly in the oxidation of alkanes and alcohols [39,40,41,42,43,44,45,46,47,48,49,50,51,52,53], the use of such compounds in catalysis is still an understudied area. In pursuit of our interest in the transition metal-catalysed peroxidative [by tert-butyl hydroperoxide (TBHP) or H2O2] oxidation of hydrocarbons in ILs [36,37,38,39], we have tested a pair of already known [51] copper(II) complexes, viz. [CuL(H2O)2]∙H2O (1) and [CuL(dea)] (2), as catalysts for cyclohexane conversion into cyclohexanol and cyclohexanone (Scheme 1), in acetonitrile or in any of the ionic liquids [bmim][NTf2] or [hmim][NTf2] as solvents and under low-power (10 W) microwave irradiation. The favorable effect of microwaves (MWs) in comparison with conventional heating has been recognized in alkane oxidation [54].
The application of ILs in alkane functionalization is still in its infancy, in spite of their attractiveness as a green alternative to volatile organic solvents. Another eventual advantage in using IL is the possibility of recycling [39], which is not the case for the use of conventional solvents. In this work, the imidazolium-based ILs mentioned above are tested in order to try to fill this gap.
Thus, in this study, we intend to achieve to following aims: (i) to apply the known copper(II) complexes based on 2-(2-(1,3-dioxo-1H-inden-2(3H)-ylidene)hydrazinyl)benzenesulfonate as catalysts for the peroxidative oxidation of cyclohexane, in acetonitrile and in an IL; (ii) to compare the oxidation performance of H2O2 and TBHP; and (iii) to recycle the IL + catalyst system after the reaction in order to test its stability.

2. Results and Discussion

2.1. Conventional Medium

Table 1 and Table 2 present the results for the peroxidative oxidation of cyclohexane in acetonitrile, using 1 and 2 as catalysts, respectively, and also comprising the studies of the effect of the catalyst and peroxide amounts, type of peroxide and reaction time.
The product yield of the microwave assisted the peroxidative oxidation of cyclohexane depends on the catalyst amount, on the time of reaction, and on the type of oxidant, as expected. For example, the yield of cyclohexanol + cyclohexanone increases when TBHP is used instead of H2O2 (Table 1, e.g., compare entries 7 and 15), and the tested catalyst amounts (10−4 and 10−5 mol·L−1) when using TBHP also increases with the decrease of the amount of catalyst 1, probably due to a more extensive overoxidation when using the higher catalyst amount. Upon increasing the reaction time, the yield of product also increases at least up to a maximum of 2 h reaction time. For this reaction time of 2 h and using 1 × 10−5 mol·L−1 of catalyst 1, 29.6% yield of total product was obtained using TBHP as oxidant (Table 1, entry 15).
When using H2O2 as oxidant, a higher selectivity towards the alcohol is observed, while with TBHP, the amount of this product is lower than that of the ketone (Table 1, entries 9–12), and eventually no alcohol is detected (Table 1, entries 13–16).
It was also observed that with TBHP and a concentration of catalyst 1 of 1 × 10−5 mol·L−1, TONs above 6 × 103 are attained, reaching 136 × 102 in 2 h (Table 1, entry 15). A TOF value above 120 × 102 h−1 is achieved at the shorter reaction time of 30 min (Table 1, entry 13).
In comparison to 1, the catalytic activity of 2 is lower (Table 2). However, the yields of the product almost always increase with time and for both types of oxidant. In addition, catalyst 2 is 100% selective for the ketone versus the alcohol by using TBHP as oxidant (Table 2, entries 9–16). The maximum TON value achieved with 2 using TBHP and the lowest tested concentration of the catalyst, 708 × 10 in 2 h (Table 2, entry 15), is nearly half that achieved with 1 under the same experimental conditions (136 × 102, Table 1, entry 15).
The catalytic performances of 1 are comparable in terms of yield with those obtained with aqua complexes of iron with arylhydrazone-β-diketone ligands, at room temperature, but using the higher catalyst loading of ca.1 × 10−3 mol·L−1, the 1:7.5 substrate/oxidant ratio and 6 h reaction, achieving TON values not higher than 290 [52]. With catalyst 1, a maximum TON value of 13 × 103 was attained using 1 × 10−5 mol·L−1 of catalyst 1 and only the 1:2 substrate/oxidant ratio. Similar yields were achieved with copper complexes [55], but for 6 h reaction and the substrate/oxidant ratio of 1:10.

2.2. Unconventional Medium (Ionic Liquid)

In view of the better catalytic performance of 1 relative to 2, the former catalyst, at the lowest concentration of 1 × 10−5 mol·L−1, was used for the peroxidative oxidation of cyclohexane and using the ionic liquid [bmim][NTf2] or [hmim][NTf2] as solvent. Hydrogen peroxide was kept as the oxidant as it is inexpensive and environmentally friendly. No acidic additive was used. Higher yields of product were achieved in [bmim][NTf2], reaching 13.9% in 3 h (Table 3, entry 4; compare with Table 1, entry 8). After such a time of reaction, the cyclohexanol/cyclohexanone ratio is similar to that obtained with acetonitrile as solvent, but the TON value is now higher (639 × 10 against 386 × 10; compare Table 3, entry 4 and entry 8 in Table 1). That ratio reaches a maximum of 2.4 upon 2 h reaction, with cyclohexanol being here the major product; the alcohol yield then decreases with a longer reaction time, while the ketone yield increases (Table 3; compare entries 3 and 4).
By using [hmim][NTf2] as solvent, a very low yield of 0.34% was obtained after 3 h (Table 3, entry 8), much lower than those observed for [bmim][NTf2] and even acetonitrile (13.9%, Table 3, entry 4; 8.4%, Table 1, entry 8). This difference in yield may eventually be due to the higher viscosity of [hmim][NTf2] relative to [bmim][NTf2] (26.2 mPa·s for the former and 20.4 mPa·s for the latter, at 323.15 K) [56].
The performance of catalyst 1 after recycling is presented in Table 4, which clearly indicates a severe loss of activity.
For homogeneous catalysis, the main industrial process exhibits yields of 4–10% of KA (ketone + alcohol) oil (cyclohexanol + cyclohexanone) with a maximum 85% selectivity at 150 °C [57]. The present study shows a visible improvement, both in yield and in using milder conditions. Other homogeneous Cu(II) complexes of arylhydrazone published previously led to overall yields up to 34% and TONs up to 42 [58]. For Cu(II) complexes of pyrazole the obtained yields were up to 58% and TONs up to 108 × 10 in 30 min reaction time at 100 °C [59]. In our study, our TONs are higher.
The peroxidative oxidation of cyclohexane catalyzed by polynuclear copper(II) complexes of arylhydrazone with H2O2, was also carried out in [bmim][BF4], in which a yield of 29.5% was achieved, at 90 min reaction and the same temperature as used here, but with 10−3 mol·L−1 of catalyst [39], which is two orders of magnitude higher than that used in our work.
Regarding the effect of the IL as the reaction medium, we note that the interaction of the scorpionate iron(II) catalyst [FeCl2(Tpm)] (Tpm = hydrotris(pyrazol-1-yl)methane) with the IL 1-butyl-3-methylimidazolium dicyanamide ([bmim][N(CN)2]) was investigated using theoretical DFT calculations [28]. They indicate the coordination of the dicyanamide anion to form the neutral associate [bmim][FeCl2{N(CN)2}(Tpm)], which accounts for the effective retention of the catalyst by the IL, without appreciable loss upon recycling. In the current case of the copper(II) complexes [CuL(H2O)2]∙H2O (1) and [CuL(dea)] (2), the interaction of the copper sites with the NTf2- anion of the IL is expected to be weaker for steric reasons in view of the bulkiness of this anion, which does not allow the prevention of catalyst leaching. This can account, at least in part, for the ineffective catalyst recycling, as it is shown in Figure 1. The possible formation of inactive copper compounds (e.g., copper oxido species) can also contribute to such behaviour.

3. Experimental Section

3.1. Materials and Equipment

All chemicals were obtained from commercial sources and used as received. All the catalytic work was performed in a Microwave Synthesis Reactor (Anton Paar, Graz, Austria) at 50 °C. Chromatographic analyses were undertaken by using a Fisons Instruments GC 8000 (Fisons Instruments SpA, Rodano, Italy) series gas chromatograph with a DB-624 (J&W) capillary column (Agilent, Santa Clara, CA, USA) (flame ionization detector) (temperature of injector: 230 °C) and the Jasco-Borwin v.1.50 software (ChromatographyForum, Ashford, UK) (temperature range 100–180 °C). The internal standard method was used to quantify the organic products (see below for details).

3.2. Synthesis of the Catalysts

The synthesis and characterization of both copper(II) catalysts (Scheme 2) were reported earlier [51] and are therefore not discussed here.

3.3. Oxidation of Cyclohexane and Products Analysis

Cyclohexane oxidations were carried out in a microwave (MW) reactor. Typically, 3 mL of acetonitrile was used (in the experiments with this solvent), to which the solid catalyst was added, and the solution was stirred to ensure the total dissolution of the catalyst. Then, the substrate (cyclohexane) was added to the reaction mixture (initial cyclohexane concentration in the reaction solution of 0.46 mol L−1), followed by the acid additive (concentrated H2SO4, 5 μL, 0.094 mmol) and the reaction started upon the addition of the oxidant (aq. 30% or 50% H2O2 or aq. 70% TBHP) in one portion (initial concentration in the reaction solution of 0.92 mol L−1). Other acidic additives were tested, such as HNO3, but no product was detected. The reaction was subjected to microwave irradiation (10 W) for variable periods of time and at a temperature of 50 °C (for comparison with other works, which have also used this temperature).
In the experiments with the ionic liquid ([bmim][NTf2] or [hmim][NTf2]), 2 mL of IL was used instead of NCMe, and the same reaction conditions mentioned above were applied, except for the fact that, in this case, no acidic additive was used and only aq. H2O2 was applied as oxidant. After the reaction with IL, 0.500 mL of distilled water was added and the mixture stirred to ensure the extraction of the products to the aqueous phase. A further amount of distilled water was added after the analysis, but no additional product was extracted.
The products were identified by GC, by the comparison of the retention times with those of commercial products. Cyclopentanone (0.05 mL) was used as an internal standard. The amount of formed product was estimated by comparing the retention times and peak areas of the reaction species with the ones of commercially available products. An example of yield calculation is presented in Supplementary Materials. PPh3 was added to the samples, to ensure that all the cyclohexylhydroperoxide present in the mixture was converted to cyclohexanol [60,61,62]. Upon each PPh3 addition, a little effervescence occurred. When no more effervescence was observed, the addition of PPh3 was stopped.

4. Conclusions

In the present work, it was observed that catalyst 1 is more active than 2, in accordance with its lower coordination number and the presence of two labile water ligands. Moreover, tert-butyl-hydroperoxide, a stronger oxidant than H2O2, gives rise to higher product yields, and with a higher selectivity towards cyclohexanone instead of cyclohexanol. With [bmim][NTf2], good results were achieved, but only with catalyst 1, even with a small catalyst amount and with the weaker oxidant H2O2. By increasing the cation chain length (use of [hmim][NTf2]), the product yield sharply decreases. Although the higher viscosity of the latter IL may have an effect, further studies with ILs with cations with different sizes must be undertaken to understand the effect of the IL size.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/8/12/636/s1, Figure S1. Example of GC chromatogram; Figure S2. Calibration line for cyclohexanone; Table S1. Species areas.

Author Contributions

G.A.O.T. did the synthesis of the complexes, the catalytic tests and wrote the majority of the paper; A.P.C.R. did the plan of the catalytic studies and wrote the discussion of the catalytic results; M.F.C.G.d.S. solved the complexes structures by X-ray diffraction; K.T.M. helped in the complexes synthesis and wrote a part of the introduction; L.C.B. gave general revision; A.J.L.P. did the final revision of the paper.

Funding

This research was founded by the Foundation for Science and Technology (FCT), grant PD/BD/106015/2014. The APC was funded by SFRH/BPD/90883/2012).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Labinger, J.A. Platinum-Catalyzed C-H Functionalization. Chem. Rev. 2017, 117, 8483–8496. [Google Scholar] [CrossRef]
  2. Pombeiro, A.J.L. Toward Functionalization of Alkanes Under Environmentally Benign Conditions. In Advances in Organometallic Chemistry and Catalysis: The Silver/Gold Jubilee International Conference on Organometallic Chemistry Celebratory Book; Pombeiro, A.J.L., Ed.; Wiley: Hoboken, NJ, USA, 2014; pp. 15–23. ISBN 978-1-11-851014-8. [Google Scholar]
  3. Crabtree, R.H. Alkane C-H activation and functionalization with homogeneous transition metal catalysts: A century of progress-a new millennium in prospect. J. Chem. Soc. Dalton Trans. 2001, 17, 2437–2450. [Google Scholar] [CrossRef]
  4. Crabtree, R.H.A.E. Shilov’s influence on early work in organometallic CH activation and functionalization. J. Org. Chem. 2015, 793, 41–46. [Google Scholar] [CrossRef]
  5. Shul’pin, G.B. New Trends in Oxidative Functionalization of Carbon-Hydrogen Bonds: A Review. Catalysts 2016, 6, 50. [Google Scholar] [CrossRef]
  6. Nesterov, D.S.; Nesterova, O.V.; Pombeiro, A.J.L. Homo- and heterometallic polynuclear transition metal catalysts for alkane C-H bonds oxidative functionalization: Recent advances. Coord Chem. Rev. 2018, 355, 199–222. [Google Scholar] [CrossRef]
  7. Banerjee, A.; Sarkar, S.; Patel, B.K. C-H functionalisation of cycloalkanes. Org. Biomol. Chem. 2017, 15, 505–530. [Google Scholar] [CrossRef]
  8. Schuchardt, U.; Carvalho, W.A.; Spinace, E.V. Why is it interesting to study cyclohexane oxidation. Synlett 1993, 10, 713–718. [Google Scholar] [CrossRef]
  9. Schuchardt, U.; Cardoso, D.; Sercheli, R.; Pereira, R.; de Cruz, R.S.; Guerreiro, M.C.; Mandelli, D.; Spinace, E.V.; Fires, E.L. Cyclohexane oxidation continues to be a challenge. Appl. Catal. A 2001, 211, 1–17. [Google Scholar] [CrossRef]
  10. Shilov, A.E.; Shul’pin, G.B. (Eds.) Activation and Catalytic Reactions of Saturated Hydrocarbons in the Presence of Metal Complexes; Springer Science & Business Media: New York, NY, USA, 2006; Volume 21. [Google Scholar]
  11. Wu, W.; Jiang, H. Palladium-Catalyzed Oxidation of Unsaturated Hydrocarbons Using Molecular Oxygen. Acc. Chem. Res. 2012, 45, 1736–1748. [Google Scholar] [CrossRef]
  12. Allen, S.E.; Walvoord, R.R.; Padilla-Salinas, R.; Kozlowski, M.C. Aerobic Copper-Catalyzed Organic Reactions. Chem. Rev. 2013, 113, 6234–6458. [Google Scholar] [CrossRef] [Green Version]
  13. Shul’pin, G.B.; Loginov, D.A.; Shul’pina, L.S.; Ikonnikov, N.S.; Idrisov, V.O.; Vinogradov, M.M.; Osipov, S.N.; Nelyubina, Y.V.; Tyubaeva, P.M. Stereoselective Alkane Oxidation with meta-Chloroperoxybenzoic Acid (MCPBA) Catalyzed by Organometallic Cobalt Complexes. Molecules 2016, 21, 1593. [Google Scholar] [CrossRef]
  14. Saisaha, P.; Dong, J.J.; Meinds, T.G.; de Boer, J.W.; Hage, R.; Mecozzi, F.; Kasper, J.B.; Browne, W.R. Mechanism of Alkene, Alkane, and Alcohol Oxidation with H2O2 by an in Situ Prepared Mn-II/Pyridine-2-carboxylic Acid Catalyst. ACS Catal. 2016, 6, 3486–3495. [Google Scholar] [CrossRef]
  15. Dragancea, D.; Talmaci, N.; Shova, S.; Novitchi, G.; Darvasiová, D.; Rapta, P.; Breza, M.; Galanski, M.; Kožıšek, J.; Martins, N.M.R.; et al. Vanadium(V) Complexes with Substituted 1,5-bis(2-hydroxybenzaldehyde)carbohydrazones and Their Use As Catalyst Precursors in Oxidation of Cyclohexane. Inorg. Chem. 2016, 55, 9187–9203. [Google Scholar] [CrossRef]
  16. Silva, T.F.S.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Kuznetsov, M.L.; Fernandes, A.R.; Silva, A.; Pan, C.-J.; Lee, J.-F.; Hwang, B.J.; Pombeiro, A.J.L. Cobalt Complexes with Pyrazole Ligands as Catalyst Precursors for the Peroxidative Oxidation of Cyclohexane: X-ray Absorption Spectroscopy Studies and Biological Applications. Chem. Asian J. 2014, 9, 1132–1143. [Google Scholar] [CrossRef]
  17. Martins, L.M.D.R.S.; Pombeiro, A.J.L. Tris(pyrazol-1-yl)methane metal complexes for catalytic mild oxidative functionalizations of alkanes, alkenes and ketones. Coord. Chem. Rev. 2014, 265, 74–88. [Google Scholar] [CrossRef]
  18. Hazra, S.; Mukherjee, S.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. A cyclic tetranuclear cuboid type copper(II) complex doubly supported by cyclohexane-1,4-dicarboxylate: Molecular and supramolecular structure and cyclohexane oxidation activity. RSC Adv. 2014, 4, 48449–48457. [Google Scholar] [CrossRef]
  19. Timokhin, I.; Pettinari, C.; Marchetti, F.; Pettinari, R.; Condello, F.; Galli, S.; Alegria, E.C.B.A.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Novel Coordination Polymers with (Pyrazolato)-Based Tectons: Catalytic Activity in the Peroxidative Oxidation of Alcohols and Cyclohexane. Cryst. Growth Des. 2015, 15, 2303–2317. [Google Scholar] [CrossRef]
  20. Ribeiro, A.P.C.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. N2O-Free single-pot conversion of cyclohexane to adipic acid catalysed by an iron(II) scorpionate complex. Green Chem. 2017, 19, 1499–1501. [Google Scholar] [CrossRef]
  21. Nesterova, O.V.; Nesterov, D.S.; Krogul-Sobczak, A.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Synthesis, crystal structures and catalytic activity of Cu(II) and Mn(III) Schiff base complexes: Influence of additives on the oxidation catalysis of cyclohexane and 1-phenylethanol. J. Mol. Catal. A Chem. 2017, 426, 506–515. [Google Scholar] [CrossRef]
  22. Wasserscheid, P. Transition Metal Catalysis in Ionic Liquids. In Handbook of Green Chemistry; Dupont, J., Kollar, L., Eds.; Wiley-VCH Verlag GmbH & Co.: Weinheim, Germany, 2010; Volume 6, pp. 65–91. ISBN 978-3-52-762869-8. [Google Scholar]
  23. Pârvulescu, V.I.; Hardacre, C. Catalysis in ionic liquids. Chem. Rev. 2007, 107, 2615–2665. [Google Scholar] [CrossRef]
  24. Vekariya, R.L. A review of ionic liquids: Applications towards catalytic organic transformations. J. Mol. Liq. 2017, 227, 44–60. [Google Scholar] [CrossRef]
  25. Dai, C.; Zhang, J.; Huang, C.; Lei, Z. Ionic Liquids in Selective Oxidation: Catalysts and Solvents. Chem. Rev. 2017, 117, 6929–6983. [Google Scholar] [CrossRef]
  26. Li, Z.; Xia, C.-G. Oxidation of hydrocarbons with iodobenzene diacetate catalyzed by manganese(III) porphyrins in a room temperature ionic liquid. J. Mol. Catal. A Chem. 2004, 214, 95–101. [Google Scholar] [CrossRef]
  27. Li, Z.; Xia, C.-G.; Xu, C.-Z. Oxidation of alkanes catalyzed by manganese(III) porphyrin in an ionic liquid at room temperature. Tetrahedron Lett. 2003, 44, 9229–9232. [Google Scholar] [CrossRef]
  28. Ribeiro, A.P.C.; Martins, L.M.D.R.S.; Kuznetsov, M.L.; Pombeiro, A.J.L. Tuning Cyclohexane Oxidation: Combination of Microwave Irradiation and Ionic Liquid with the C-Scorpionate [FeCl2(Tpm)] Catalyst. Organometallics 2017, 36, 192–198. [Google Scholar] [CrossRef]
  29. Martins, L.M.D.R.S.; Martins, A.; Alegria, E.C.B.A.; Carvalho, A.P.; Pombeiro, A.J.L. Efficient cyclohexane oxidation with hydrogen peroxide catalysed by a C-scorpionate iron(II) complex immobilized on desilicated MOR zeolite. Appl. Catal. A-Gen. 2013, 464, 43–50. [Google Scholar] [CrossRef] [Green Version]
  30. Silva, T.F.S.; Guedes da Silva, M.F.C.; Mishra, G.S.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Synthesis and structural characterization of iron complexes with 2,2,2-tris(1-pyrazolyl)ethanol ligands: Application in the peroxidative oxidation of cyclohexane under mild conditions. J. Org. Chem. 2011, 696, 1310–1318. [Google Scholar] [CrossRef]
  31. Silva, T.F.S.; Alegria, E.C.B.A.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Half-sandwich scorpionate vanadium, iron and copper complexes: Synthesis and application in the catalytic peroxidative oxidation of cyclohexane under mild conditions. Adv. Synth. Catal. 2008, 350, 706–716. [Google Scholar] [CrossRef]
  32. Silva, T.F.S.; Luzyanin, K.V.; Kirillova, M.V.; Guedes da Silva, M.F.C.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Novel Scorpionate and Pyrazole Dioxovanadium Complexes, Catalysts for Carboxylation and Peroxidative Oxidation of Alkanes. Adv. Synth. Catal. 2010, 352, 171–187. [Google Scholar] [CrossRef] [Green Version]
  33. Martins, L.M.D.R.S.; Pombeiro, A.J.L. Water-Soluble C-Scorpionate Complexes–Catalytic and Biological Applications. Eur. J. Inorg. Chem. 2016, 2236–2252. [Google Scholar] [CrossRef]
  34. Sutradhar, M.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.G.; Pombeiro, A.J.L. Vanadium complexes: Recent progress in oxidation catalysis. Coord. Chem. Rev. 2015, 301, 200–239. [Google Scholar] [CrossRef]
  35. Martins, L.M.D.R.S. C-Homoscorpionate Oxidation Catalysts-Electrochemical and Catalytic Activity. Catalysts 2017, 7, 12. [Google Scholar] [CrossRef]
  36. Paul, A.; Ribeiro, A.P.C.; Karmakar, A.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. A Cu(II) MOF with a flexible bifunctionalised terpyridine as an efficient catalyst for the single-pot hydrocarboxylation of cyclohexane to carboxylic acid in water/ionic liquid médium. Dalton Trans. 2016, 45, 12779–12789. [Google Scholar] [CrossRef]
  37. Hazra, S.; Ribeiro, A.P.C.; Guedes da Silva, M.F.C.; Nieto de Castro, C.A.; Pombeiro, A.J.L. Syntheses and crystal structures of benzene-sulfonate and -carboxylate copper polymers and their application in the oxidation of cyclohexane in ionic liquid under mild conditions. Dalton Trans. 2016, 45, 13957–13968. [Google Scholar] [CrossRef]
  38. Ribeiro, A.P.C.; Martins, L.M.D.R.S.; Hazra, S.; Pombeiro, A.J.L. Catalytic oxidation of cyclohexane with hydrogen peroxide and a tetracopper(II) complex in an ionic liquid. Comp. Rend. Chim. 2015, 18, 758–765. [Google Scholar] [CrossRef]
  39. Jlassi, R.; Ribeiro, A.P.C.; Guedes da Silva, M.F.C.; Mahmudov, K.T.; Kopylovich, M.N.; Anisimova, T.B.; Naïli, H.; Tiago, G.A.O.; Pombeiro, A.J.L. Polynuclear Copper(II) Complexes as Catalysts for the Peroxidative Oxidation of Cyclohexane in a Room-Temperature Ionic Liquid. Eur. J. Inorg. Chem. 2014, 2014, 4541–4550. [Google Scholar] [CrossRef]
  40. Mazzi, A.; Paul, S.; Cavani, F.; Wojcieszak, R. Cyclohexane Oxidation to Adipic Acid Under Green Conditions: A Scalable and Sustainable Process. ChemCatChem 2018, 10, 3680–3682. [Google Scholar] [CrossRef]
  41. Gryca, I.; Czerwinska, K.; Machura, B.; Chrobok, A.; Shul’pina, L.S.; Kuznetsov, M.L.; Nesterov, D.S.; Kozlov, Y.N.; Pombeiro, A.J.L.; Varyan, I.A.; et al. High Catalytic Activity of Vanadium Complexes in Alkane Oxidations with Hydrogen Peroxide: An Effect of 8-Hydroxyquinoline Derivatives as Noninnocent Ligands. Inorg. Chem. 2018, 57, 1824–1839. [Google Scholar] [CrossRef]
  42. Ribeiro, A.P.C.; Alegria, E.C.B.A.; Kopylovich, M.N.; Ferraria, A.M.; do Rego, A.M.B.; Pombeiro, A.J.L. Comparison of microwave and mechanochemical energy inputs in catalytic oxidation of cyclohexane. Dalton Trans. 2018, 47, 8193–8198. [Google Scholar] [CrossRef]
  43. Alegria, E.C.B.A.; Fontolan, E.; Ribeiro, A.P.C.; Kopylovich, M.N.; Domingos, C.; Ferraria, A.M.; Bertani, R.; do Rego, A.M.B.; Pombeiro, A.J.L. Simple solvent-free preparation of dispersed composites and their application as catalysts in oxidation and hydrocarboxylation of cyclohexane. Mater. Today Chem. 2017, 5, 52–62. [Google Scholar] [CrossRef]
  44. Aboelfetoh, E.F.; Pietschnig, R. Preparation and Catalytic Performance of Al2O3, TiO2 and SiO2 Supported Vanadium Based-Catalysts for C-H Activation. Catal. Lett. 2009, 127, 83–94. [Google Scholar] [CrossRef]
  45. Aboelfetoh, E.F.; Fechtelkord, M.; Pietschnig, R. Structure and catalytic properties of MgO-supported vanadium oxide in the selective oxidation of cyclohexane. J. Mol. Catal. A Chem. 2010, 318, 51–59. [Google Scholar] [CrossRef]
  46. Aboelfetoh, E.F.; Pietschnig, R. Preparation, Characterization and Catalytic Activity of MgO/SiO2 Supported Vanadium Oxide Based Catalysts. Catal. Lett. 2014, 144, 97–103. [Google Scholar] [CrossRef]
  47. Mahmudov, K.T.; Kopylovich, M.N.; Pombeiro, A.J.L. Coordination chemistry of arylhydrazones of methylene active compounds. Coord. Chem. Rev. 2013, 257, 1244–1281. [Google Scholar] [CrossRef]
  48. Mahmudov, K.T.; Kopylovich, M.N.; Sabbatini, A.; Drew, M.G.B.; Martins, L.M.D.R.S.; Pettinari, C.; Pombeiro, A.J.L. Cooperative Metal Ligand Assisted E/Z lsomerization and Cyano Activation at Cu-II and Co-II Complexes of Arylhydrazones of Active Methylene Nitriles. Inorg. Chem. 2014, 53, 9946–9958. [Google Scholar] [CrossRef]
  49. Mahmudov, K.T.; Pombeiro, A.J.L. Resonance-Assisted Hydrogen Bonding as a Driving Force in Synthesis and a Synthon in the Design of Materials. Chem. Eur. J. 2016, 22, 16356–16398. [Google Scholar] [CrossRef]
  50. Mahmudov, K.T.; Kopylovich, M.N.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Non-covalent interactions in the synthesis of coordination compounds: Recent advances. Coord. Chem. Rev. 2017, 345, 54–72. [Google Scholar] [CrossRef]
  51. Tiago, G.A.O.; Ribeiro, A.P.C.; Mahmudov, K.T.; Guedes da Silva, M.F.C.; Branco, L.C.; Pombeiro, A.J.L. Mononuclear copper(II) complexes of an arylhydrazone of 1H-indene-1,3(2H)-dione as catalysts for the oxidation of 1-phenylethanol in ionic liquid medium. RSC Adv. 2016, 6, 83412–83420. [Google Scholar] [CrossRef]
  52. Kopylovich, M.N.; Mac Leod, T.C.O.; Haukka, M.; Amanullayeva, G.I.; Mahmudov, K.T.; Pombeiro, A.J.L. Aquasoluble iron(III)-arylhydrazone-beta-diketone complexes: Structure and catalytic activity for the peroxidative oxidation of C-5-C-8 cycloalkanes. J. Inorg. Biochem. 2012, 115, 72–77. [Google Scholar] [CrossRef]
  53. Jlassi, R.; Ribeiro, A.P.C.; Alegria, E.C.B.A.; Naïli, H.; Tiago, G.A.O.; Rüffer, T.; Lang, H.; Zubkov, F.I.; Pombeiro, A.J.L.; Rekik, W. Copper(II) complexes with an arylhydrazone of methyl 2-cyanoacetate as effective catalysts in the microwave-assisted oxidation of cyclohexane. Inorg. Chim. Acta 2018, 471, 658–663. [Google Scholar] [CrossRef]
  54. Kustov, L.M.; Kucherov, A.V.; Finashina, E.D. Oxidative dehydrogenation of C-2-C-4 alkanes into alkenes: Conventional catalytic systems and microwave catalysis. Russ. J. Phys. Chem. A 2013, 87, 345–351. [Google Scholar] [CrossRef]
  55. Kopylovich, M.N.; Nunes, A.C.C.; Mahmudov, K.T.; Haukka, M.; Mac Leod, T.C.O.; Martins, L.M.D.R.S.; Kuznetsov, M.L.; Pombeiro, A.J.L. Complexes of copper(II) with 3-(ortho-substituted phenylhydrazo)pentane-2,4-diones: Syntheses, properties and catalytic activity for cyclohexane oxidation. Dalton Trans. 2011, 40, 2822–2836. [Google Scholar] [CrossRef]
  56. Tariq, M.; Carvalho, P.J.; Coutinho, J.A.P.; Marrucho, I.M.; Canongia Lopes, J.N.C.; Rebelo, L.P.N. Viscosity of (C-2-C-14) 1-alkyl-3-methylimidazolium bis(trifluoromethylsulfonyl)amide ionic liquids in an extended temperature range. Fluid Phase Equilib. 2011, 301, 22–32. [Google Scholar] [CrossRef]
  57. Weissermel, K.; Arpe, H.-J. Industrial Organic Chemistry, 2nd ed.; VCH Press: Weinheim, Germany, 1993; ISBN 978-3-52-761919-1. [Google Scholar]
  58. Di Nicola, C.; Garau, F.; Karabach, Y.Y.; Martins, L.M.D.R.S.; Monari, M.; Pandolfo, L.; Pettinari, C.; Pombeiro, A.J.L. Trinuclear Triangular Copper(II) Clusters. Synthesis, Electrochemical Studies and Catalytic Peroxidative Oxidation of Cycloalkanes. Eur. J. Inorg. Chem. 2009, 666–676. [Google Scholar] [CrossRef]
  59. Galassi, R.; Simon, O.C.; Burini, A.; Tosi, G.; Conti, C.; Graiff, C.; Martins, N.M.R.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L.; Martins, L.M.D.R.S. Copper(I) and copper(II) metallacycles as catalysts for microwave assisted selective oxidation of cyclohexane. Polyhedron 2017, 134, 143–152. [Google Scholar] [CrossRef]
  60. Shul’pin, G.B.; Attanasio, D.; Suber, L. Oxidations by a H2O2-VO3- pyrazine-2-carboxylic acid reagent. Russ. Chem. Bull. 1993, 42, 55–59. [Google Scholar] [CrossRef]
  61. Nizova, G.V.; Süss-Fink, G.; Shul’pin, G.B. Oxidations by the reagent O2–H2O2–vanadium complex–pyrazine-2-carboxylic acid: Efficient oxygenation of methane and other lower alkanes in acetonitrile. Tetrahedron 1997, 53, 3603–3614. [Google Scholar] [CrossRef]
  62. Guerreiro, M.C.; Schuchardt, U.; Shul’pin, G.B. Oxidations with the “O2-H2O2-VO3- pyrazine-2-carboxylic acid” reagent. Russ. Chem. Bull. 1997, 46, 749–754. [Google Scholar] [CrossRef]
Scheme 1. Catalytic oxidation of cyclohexane. IL: ionic liquids; MW: microwave.
Scheme 1. Catalytic oxidation of cyclohexane. IL: ionic liquids; MW: microwave.
Catalysts 08 00636 sch001
Figure 1. Recycling of the IL + catalyst 1 system.
Figure 1. Recycling of the IL + catalyst 1 system.
Catalysts 08 00636 g001
Scheme 2. Copper catalysts used in this work [51].
Scheme 2. Copper catalysts used in this work [51].
Catalysts 08 00636 sch002
Table 1. Oxidation of cyclohexane to cyclohexanol and cyclohexanone catalysed by 1 a.
Table 1. Oxidation of cyclohexane to cyclohexanol and cyclohexanone catalysed by 1 a.
EntryOxidantCatalyst Amount
(mol L−1)
Time (min)Yield % (after PPh3) bTotal TON dTotal TOF e
(h−1)
[Alc]/[Keto] f
Cyclo-HexanoneCyclo-HexanolTotal c
1H2O2 1 × 10−4300.84.25.02304605.3
2603.85.08.94094091.3
31204.45.29.64422211.2
41804.78.112.85891961.7
5H2O2 1 × 10−5302.33.25.4248 × 10497 × 101.4
6602.24.56.7308 × 10308 × 102.0
71204.44.79.1419 × 10209 × 101.1
81803.74.78.4386 × 10129 × 101.3
9TBHP 1 × 10−4304.71.76.42945880.4
10608.42.210.64884880.3
1112013.02.315.37043520.2
1218021.72.824.6113 × 103770.1
13TBHP1 × 10−53013.50.013.5621 × 10124 × 1020.0
146020.20.020.2929 × 10929 × 100.0
1512029.60.029.6136 × 102681 × 100.0
1618017.30.017.3796 × 10265 × 100.0
17 gH2O2 1 × 10−5120trace0.00.0---
18 gTBHP 1 × 10−5120trace0.00.0---
19 hTBHP 4 × 10−41201.63.85.449242.4
a Reaction conditions, unless stated otherwise: [cyclohexane]0 = 0.46 mol L−1, [oxidant]0 = 0.92 mol L−1, additive (H2SO4) = 0.019 mol L−1, CH3CN (3 mL), 50 °C. b Based on GC (Gas Chromatography) analysis, after treatment with PPh3. c Values correspond to total yields (moles of products/100 moles of cyclohexane). d TON = Total turnover number (moles of product/mol of catalyst). e TOF (h-1) = turnover frequency (TON/time). f Ratio between the concentrations of cyclohexanol (Alc) and cyclohexanone (Keto). g Blank test (no metal catalyst). h Cu(NO3)2 as catalyst [55].
Table 2. Oxidation of cyclohexane to cyclohexanol and cyclohexanone catalysed by 2 a.
Table 2. Oxidation of cyclohexane to cyclohexanol and cyclohexanone catalysed by 2 a.
EntryOxidantCatalyst Amount
(mol L−1)
Time (min)Yield (%) (after PPh3) bTotal TON dTotal TOF e
(h−1)
[Alc]/[Keto] f
Cyclo-HexanoneCyclo-HexanolTotal c
1H2O2 1 × 10−4301.10.31.4641280.3
2601.00.51.569690.5
31201.40.92.3106530.6
41802.00.82.8129430.4
5H2O2 1 × 10−5301.20.31.5690138 × 100.3
6603.51.75.2239 × 10239 × 100.5
71203.92.26.1281 × 10140 × 100.6
81805.43.58.9409 × 10137 × 100.6
9TBHP 1 × 10−43017.40.017.4800160 × 100.0
106018.40.018.48468460.0
1112020.00.020.09204600.0
1218024.50.024.5113 × 103760.0
13TBHP 1 × 10−5307.70.07.7354 × 10708 × 100.0
146015.20.015.2699 × 10699 × 100.0
1512015.40.015.4708 × 10354 × 100.0
1618011.10.011.1511 × 10170 × 100.0
17 gH2O21 × 10−5120trace0.00.0---
18 gTBHP1 × 10−5120trace0.00.0---
a Reaction conditions, unless stated otherwise: [cyclohexane]0 = 0.46 mol L−1, [oxidant]0 = 0.92 mol L−1, additive (H2SO4) = 0.019 mol L−1, CH3CN (3 mL), 50 °C. b Based on GC analysis, after treatment with PPh3. c Values correspond to total yields (moles of products/100 moles of cyclohexane). d TON = Total turnover number (moles of product/mol of catalyst). e TOF (h-1) = turnover frequency (TON/time). f Ratio between the concentrations of cyclohexanol (Alc) and cyclohexanone (Keto). g Blank test (no metal catalyst).
Table 3. Oxidation of cyclohexane to cyclohexanol and cyclohexanone catalysed by 1 a in an IL.
Table 3. Oxidation of cyclohexane to cyclohexanol and cyclohexanone catalysed by 1 a in an IL.
EntryOxidantTime (min)Yield (%) (after PPh3) bTotal TON dTotal TOF e
(h−1)
[Alc]/[Keto] f
CyclohexanoneCyclohexanolTotal c
[bmim][NTf2]
1H2O2 303.65.28.8405 × 10810 × 101.4
2604.45.910.3474 × 10474 × 101.3
31203.58.512.0552 × 10276 × 102.4
41806.17.813.9639 × 10213 × 101.3
[hmim][NTf2]
5H2O2 300.00.00.000n.d.
6600.00.00.000n.d.
71200.110.100.2197490.9
81800.180.160.34156520.9
a Reaction conditions, unless stated otherwise: [cyclohexane]0 = 0.46 mol L−1, [H2O2]0 = 0.92 mol L−1, [cat] = 1 × 10−5 mol L−1, IL (2 mL), 50 °C. b Based on GC analysis, after treatment with PPh3. c Values correspond to total yields (moles of products/100 moles of cyclohexane). d TON = Total turnover number (moles of product/mol of catalyst). e TOF (h-1) = turnover frequency (TON/time). f Ratio between the concentrations of cyclohexanol (Alc) and cyclohexanone (Keto).
Table 4. Recycling of the system 1 + [bmim][NTf2] a.
Table 4. Recycling of the system 1 + [bmim][NTf2] a.
EntryYield (%) (after PPh3) bActivity (%)Total TON dTotal TOF (h−1) e[Alc]/[Keto] f
CyclohexanoneCyclohexanolTotal c
16.17.813.9100639 × 10213 × 101.3
20.30.81.186382132.7
30.50.30.864641550.6
a Reaction conditions, unless stated otherwise: [cyclohexane]0 = 0.58 mol L−1, [H2O2]0 = 1.15 mol L−1, [cat] = 1 × 10−5 mol L−1, IL (2 mL), 50 °C, reaction time = 180 min. b Based on GC analysis, after treatment with PPh3. c Values correspond to total yields (moles of products/100 moles of cyclohexane). d TON = Total turnover number (moles of product/mol of catalyst). e TOF (h-1) = turnover frequency (TON/time). f Ratio between the concentrations of cyclohexanol (Alc) and cyclohexanone (Keto).

Share and Cite

MDPI and ACS Style

Tiago, G.A.O.; Ribeiro, A.P.C.; C. Guedes da Silva, M.F.; Mahmudov, K.T.; Branco, L.C.; Pombeiro, A.J.L. Copper(II) Complexes of Arylhydrazone of 1H-Indene-1,3(2H)-dione as Catalysts for the Oxidation of Cyclohexane in Ionic Liquids. Catalysts 2018, 8, 636. https://doi.org/10.3390/catal8120636

AMA Style

Tiago GAO, Ribeiro APC, C. Guedes da Silva MF, Mahmudov KT, Branco LC, Pombeiro AJL. Copper(II) Complexes of Arylhydrazone of 1H-Indene-1,3(2H)-dione as Catalysts for the Oxidation of Cyclohexane in Ionic Liquids. Catalysts. 2018; 8(12):636. https://doi.org/10.3390/catal8120636

Chicago/Turabian Style

Tiago, Gonçalo A. O., Ana P. C. Ribeiro, M. Fátima C. Guedes da Silva, Kamran T. Mahmudov, Luís C. Branco, and Armando J. L. Pombeiro. 2018. "Copper(II) Complexes of Arylhydrazone of 1H-Indene-1,3(2H)-dione as Catalysts for the Oxidation of Cyclohexane in Ionic Liquids" Catalysts 8, no. 12: 636. https://doi.org/10.3390/catal8120636

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