Next Article in Journal
Recent Advances in Photocatalytic CO2 Utilisation Over Multifunctional Metal–Organic Frameworks
Next Article in Special Issue
Metal-Based Catalysts in Organic Synthesis
Previous Article in Journal
Editorial Special Issue “Heterogeneous Catalytic Materials: Synthesis, Characterization and Applications for Energetic Purposes”
Previous Article in Special Issue
Pd Nanoparticles and Mixture of CO2/CO/O2 Applied in the Carbonylation of Aniline
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fe(III) Complexes in Cyclohexane Oxidation: Comparison of Catalytic Activities under Different Energy Stimuli

by
Tannistha Roy Barman
1,
Manas Sutradhar
1,*,
Elisabete C. B. A. Alegria
1,2,*,
Maria de Fátima C. Guedes da Silva
1 and
Armando J. L. Pombeiro
1,*
1
Centro de Química Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisboa, Portugal
2
Departamento de Engenharia Química, Instituto Superior de Engenharia de Lisboa, Instituto Politécnico de Lisboa, R. Conselheiro Emídio Navarro, 1, 1959-007 Lisboa, Portugal
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(10), 1175; https://doi.org/10.3390/catal10101175
Submission received: 25 September 2020 / Revised: 8 October 2020 / Accepted: 9 October 2020 / Published: 13 October 2020
(This article belongs to the Special Issue Metal-Based Catalysts in Organic Synthesis)

Abstract

:
In this study, the mononuclear Fe(III) complex [Fe(HL)(NO3)(H2O)2]NO3 (1) derived from -acetylpyrazine-2-carbohydrazide (H2L) was synthesized and characterized by several physicochemical methods, e.g., elemental analysis, infrared (IR) spectroscopy, electrospray ionization mass spectrometry (ESI-MS), and single crystal X-ray diffraction analysis. The catalytic performances of 1 and the previously reported complexes [Fe(HL)Cl2] (2) and [Fe(HL)Cl(μ-OMe)]2 (3) towards the peroxidative oxidation of cyclohexane under three different energy stimuli (microwave irradiation, ultrasound, and conventional heating) were compared. 1-3 displayed homogeneous catalytic activity, leading to the formation of cyclohexanol and cyclohexanone as final products, with a high selectivity for the alcohol (up to 95%). Complex 1 exhibited the highest catalytic activity, with a total product yield of 38% (cyclohexanol + cyclohexanone) under optimized microwave-assisted conditions.

Graphical Abstract

1. Introduction

The functionalization of alkanes should provide convenient methods for building a range of valuable organic products [1,2,3,4,5,6,7,8,9]. However, due to their inertness, such processes are difficult and, e.g., the catalytic oxidation of alkanes remains one of the most challenging fields of chemistry.
The selective oxidation of cyclohexane to cyclohexanol and cyclohexanone has a significant relevance in terms of industrial and economical viewpoints, because these products are used as precursors of adipic acid which, among other applications, is used in the manufacturing of Nylon-6,6 [1,2,3,4,5,6,7,8,9,10,11,12,13,14]. In industry, that reaction is currently performed using a cobalt(II) naphthenate catalyst [15] at 160 °C and 15 bar with very low yields (4%), in order to achieve a good selectivity (80%) [1,2,9,14,15]. The search for more efficient catalytic systems through the development of novel metal-based catalysts and single-pot methodologies for the mild oxidation of alkanes is thus a challenging task [1,2,3,7,16,17,18,19,20,21,22,23], involving new catalysts with cheap and abundant metals, namely copper [24,25,26,27] and iron [28,29,30,31,32,33]. Microwave irradiation has also been shown to assist with the reaction in various cases [28,29,30,31,32,33].
In continuation of our work on the catalytic oxidation of alkanes and related species with metal-hydrazone catalysts [24,25,26,27,34,35,36,37,38,39,40,41,42,43,44], herein, we report the synthesis and characterization of the new mononuclear Fe(III) complex [Fe(HL)(NO3)(H2O)2]NO3 (1) obtained from -acetylpyrazine-2-carbohydrazide (H2L), and a comparison of its catalytic activity with those of the related compounds [Fe(HL)Cl2] (2) and [Fe(HL)Cl(μ-OMe)]2 (3) [45]. The catalytic activity of these complexes in the oxidation of cyclohexane was evaluated by conventional heating and under microwave (MW) or ultrasound (US) irradiations. MW and US irradiations were applied to activate the proposed catalytic system, eventually enhancing the yield, selectivity, and/or rate in comparison with conventional heating [46,47], but their application in alkane functionalization is still rather underexplored [48,49,50].

2. Results and Discussion

The reaction between Fe(NO3)3·9H2O and the pro-ligand -acetylpyrazine-2-carbohydrazide (H2L) in methanol results in the formation of the mononuclear Fe(III) complex [Fe(HL)(NO3)(H2O)2]NO3 (1) (Scheme 1). Complexes 2 and 3 were synthesized according to the literature [45]. These complexes were characterized by elemental analysis, infrared (IR) spectroscopy, and electrospray ionization mass spectrometry (ESI-MS) and, for 1, also by X-ray single crystal diffraction analysis. The expected characteristic stretching bands of HL in 1 are shifted to lower wavenumbers relative to those in the pro-ligand (see the experimental section), specifically the ν(NH) at 3136 cm−1 and the ν(C=O) detected at 1678 and 1646 cm−1; the band for nitrate ions occurs at the typical value of 1383 cm−1 [51]. The molecular ion was not observed in the ESI-MS analysis (see the experimental section), but instead, the species [M-(NO3)]+ resulting from the loss of the nitrate counter-ion was detected.

2.1. Crystal Structure

Crystals of 1 (Figure 1) were obtained from methanol upon slow evaporation, at room temperature. A summary of the crystallographic data and processing parameters is presented in Table 1. Complex 1 (Supplementary Materials: CCDC number 2019464) crystallized in the monoclinic P21/c space group, with the asymmetric unit containing an iron cation with -acetylpyrazine-2-carbohydrazide acting as a mononegative NpyrazineNamidoOketo chelate ligand and a OO’-donor nitrate anion sharing the equatorial binding region. The axial sites were engaged with two water ligands, and a cationic pentagonal bipyramidal iron complex was thus formed, with its charge being balanced by a nitrate counter anion. The unit cell of 1 contains five non-coordinated and disordered water molecules. The NN’O coordination mode of HL was observed in other cases [45,51] and differs from that found in complexes with ligands derived from N-acetylsalicylhydrazide [34,52,53]. The HL ligand in 1 is slightly twisted, as measured by the distance between the least-square plane defined by the pyrazine ring and the Cmethyl atom (0.539 Å); in 3, that distance is shorter (0.388 Å) and in 2, such moieties are coplanar [45]. Due to the water ligands and non-coordinated water molecules, compound 1 is involved in extensive H-bond interactions, which extend the structure to the third dimension. This contrasts with what was found in 2 and 3, resulting from the kind of contact [45], with the former being assembled in dimers and the latter giving rise to 1D chains.
The catalytic properties of 1-3 were investigated and compared for the peroxidative oxidation of cyclohexane, and the effect of the type of activating energy input on the catalytic output was also studied.

2.2. Peroxidative Oxidation of Cyclohexane

Using aqueous tert-butylhydroperoxide (t-BuOOH, 70% aqueous solution) as an oxidant in acetonitrile (NCMe) at 50 °C, the homogeneous 1-3 were screened for the microwave-assisted (MW) oxidation of cyclohexane (CyH) to cyclohexanone (Cy=O) and cyclohexanol (CyOH) as the final products (Scheme 2). The detection of no other products by Gas Chromatography–Mass Spectrometry (GC-MS) analysis suggests that the catalytic oxidation is very selective. The mechanism involved in the cyclohexane oxidation, involving the generation of cyclohexyl hydroperoxide (CyOOH) as a primary product, was corroborated by the method proposed by Shul’pin [54] and the reaction mixture injected in the gas chromatograph before and after treatment with triphenylphosphine (PPh3). The amount of CyOH had increased significantly after the addition of PPh3 to the reaction mixture, due to the reduction of CyOOH to CyOH and formation of phosphine oxide (PPh3O) (Scheme 2). The results are reported in Table 2 (yield values refer to the samples after treatment with an excess of PPh3).
In the presence of 1 and after 3 h under MW-irradiation at 50 °C, 14.9% of cyclohexane was converted into ketone-alcohol (KA) oil (cyclohexanol and cyclohexanone mixture), with cyclohexanol as the major product (selectivity up to 86% relative to KA oil after treatment with PPh3), in the absence of any co-catalyst (entry 5, Table 2), with a turnover number of 75 per Fe atom. There was always an increase in the amount of alcohol when the reaction mixture was analyzed after the addition of triphenylphosphine with respect to the existing amount before this treatment. This fact suggests the formation of cyclohexyl hydroperoxide (CyOOH) as a primary product which, after treatment with PPh3, is reduced to cyclohexanol (CyOH). As an example, entries 4 and 5 from Table 2 are presented, and from an initial CyOH yield of 7.5% (before PPh3 treatment), an increment to 13.2% was detected (after PPh3 treatment). Furthermore, a decrease in the amount of cyclohexanone (Cy=O) was observed, although with a less pronounced difference. For iron complexes 2 and 3, and for the same reaction conditions, total yields of 10.4% and 14.4% were obtained, with selectivities of 88% and 85% for the alcohol, respectively (Table 2, entries 22 and 25).
The oxidation reaction of cyclohexane was attempted under solvent-free conditions (in the absence of acetonitrile) in the presence of 1 (Table 2, entry 3), without success (the yield did not go beyond 0.74%, despite the high selectivity for the alcohol of 95%). The replacement of compounds 1-3 by their precursor salts, Fe(NO3)3.9H2O and FeCl2, resulted in much lower conversions, with total yields not exceeding 5% (Table 2, entries 28 and 29, respectively), denoting the significance of the NN’O-donor ligand -acetylpyrazine-2-carbohydrazide (HL) in the Fe coordination sphere in the promotion of the catalytic performance of complexes 1-3. The free pro-ligand did not exhibit any activity. Blank tests were performed in the absence of any of the Fe(III) compounds 1-3, but no noteworthy conversion was observed.
Complex 1 was chosen for optimizing the reaction conditions and several parameters were explored. The effect of the reaction time is shown in Figure 2a, for the period between 0.5 and 6 h, without any additive and at 50 °C (Table 2, entries 1, 2, 5, and 6). There was a gradual increase over time in the amount of KA oil formed, achieving a total yield of ca. 15 and 17% after 3 and 6 h, respectively. During the first hour, the formation of cyclohexanone was not significant (94 and 91% of selectivity for the alcohol, after 0.5 and 1 h reaction, respectively). The yield of cyclohexanone increased (from 0.1 to 1.7%, after 0.5 and 3 h, respectively) and the selectively to the alcohol decreased to 86%, possibly due to the partial oxidation of cyclohexanol.
The amount of catalyst varied between 2.5 and 20 μmol (Figure 2b), and a maximum yield of 19.4% of KA oil was verified for an amount of 20 μmol of 1 (entry 11, Table 2). This parameter does not affect only the yield, which increases from 3.8 to 19.4% with the indicated range amount of catalyst, but also the selectivity which, for cyclohexanol, decreases from 95 to 80% (entries 1 and 9-1, Table 2). In this way, we can conclude that both the increase of the reaction time and the increase in the amount of catalyst favor the formation of ketone.
When performing the oxidation reactions under the typical conditions (3 h, 50 °C, and 10 μmol of 1), the variation in the oxidant/substrate molar ratio also has an important effect on the oxidized product yields (Table 2, entries 5, 7, and 8). The yield of KA oil increased from 8.7 to 14.9 and afterwards to 21.5% when the oxidant/substrate molar ratio was changed from 1:1 to 2:1 and then to 4:1, respectively.
The effect of various additives on the peroxidative microwave-assisted oxidation of cyclohexane was also investigated. In the presence of nitric acid (HNO3), complex 1 exhibited a highly promising effect. The n(HNO3)/nʹ(catalyst) molar ratio changed from 5 to 50 (Figure 2c). The total yield increased to 37.7%, for n(HNO3)/nʹ(catalyst 1) = 25, relative to 14.9% obtained in the absence of any additive (Table 2, entries 5 and 14, respectively). Going beyond this molar ratio does not lead to a significant yield change (37.2% for n/n’ = 50) (Table 2, entries 14 and 15).
To investigate the effect of the acid additive over time, catalytic oxidation was performed in the presence of this acid additive (n/nʹ = 25) for several time periods (1, 3, 6, and 9 h). Firstly, for both cases, in the absence and presence of HNO3 (n/nʹ = 25) (Figure 2a,d), the quantity of oxygenated products (CyOH + Cy=O) practically reached the maximum after 3 h and then seemed to stabilize.
The positive effect of nitric acid has been observed for other catalytic systems involving the oxidative transformation of alkanes [26,27,55,56,57]. The presence of a certain amount of acid can promote the catalytic process, either by catalyst activation through the protonation of ligands and unsaturation of the metal center, or by promoting the properties of the oxidant.
The presence of HNO3 affects, apart from the total yield, the product distribution. In the case of 1, the selectivity for the alcohol was lower in the presence of acid, with the prolongation of time accentuating this effect. In the presence of acid, there was a clear preferential formation of CyOH (selectivity of 85%) in the first hour; a decrease of CyOH selectivity accompanied by an increase of Cy=O selectivity in the 1-3 h period; and beyond 3 h, the ratio between both products seemed to stabilize (Figure 2d).
In the presence of 1, the influences of the 2-pyrazine carboxylic acid (HPCA), trifluroacetic acid (TFA), and stable free radical 2,2,6,6-tetramethylpiperidin-l-oxyl (TEMPO) were also explored (Figure 3). After 3 h at 50 °C under MW-irradiation, the total yield of products dropped in the presence of HPCA (8.8%) and TEMPO (drastically to 1.7%), whereas in the presence of TFA, it increased from 14.9% to 28.3% (Table 2, entries 14, 19, 20, and 21), although not as effectively as for HNO3.
The effects of the presence of HNO3 (n(HNO3)/nʹ(catalyst 2 or 3) = 25) and TEMPO additives were also analyzed for compounds 2 and 3. In the case of the acid additive, although an increase in the amount of oxygenated products was observed, this effect was not so accentuated for these catalysts as for 1 (Table 2, entries 23 and 26, for 2 and 3, respectively). The presence of the TEMPO radical resulted, for both catalytic systems, as in the case of 1, in a drastic decrease in the yields (Table 2, entries 24 and 27, for 2 and 3, respectively).
The peroxidative oxidation of cyclohexane was also performed, for comparative purposes, using different types of energy inputs, apart from microwaves, namely conventional heating and ultrasound (US) irradiation. Reactions were performed for compounds 1-3, during 3 h at 50 °C and in the presence of HNO3 as an additive (n(HNO3)/nʹ(catalyst) = 25), and in the case of compound 1, for different periods of time.
If we consider the period of 3 h, compounds 1-3 responded differently to the different energy stimuli. It can be observed that 1 exhibits a better performance when the reaction is promoted by microwave radiation (Table 3, entry 8), conceivably due to its ionic character and larger dipole, which promote microwave energy absorption [31,48,58,59,60]. Accordingly, the effect of MW irradiation (in comparison with conventional heating) in the case of catalyst 3, with a symmetrical apolar molecule, is negligible (Table 3, entries 17 and 18). In this case, a different driving force, i.e., acoustic cavitation in sonochemistry, shows a more effective role (Table 3, entry 19), which is consistent with the known ultrasonic cleavage of a metal-ligand bond [61]. For the dinuclear catalyst 3, this can lead to the formation of more active mononuclear catalytic species. The reaction catalyzed by 2 does not seem to be favored by any of the radiations (MW and US), which may be due to its possible decomposition into less active or inactive species when it is under these energy inputs. It is also noteworthy to mention that, for the same period, the selectivity does not vary much when we compare the catalytic activity of 1-3 under the effect of different energy inputs (Table 3). For example, for the period of 3 h, and in the presence of 1, the selectivity for cyclohexanol varies between 66 and 69% for different energy sources. In the case of 2 and 3, only the MW stands out and the selectivities reach values higher than 90%.
In addition, compound 1 was exposed to different types of energy input for different reaction times and, in all of the cases, a maximum selectivity for cyclohexanol was observed in the first 30 min of the reaction (Figure 4, Table 3). Thereafter, the selectivity decreased, conceivably due to the conversion of the cyclohexanol into cyclohexanone.
The present Fe(III) catalytic system is more effective for the oxidation of cyclohexane under MW conditions in terms of yields, compared to Cu(II) or V(V) catalytic systems with hydrazone-based ligands [27,35,42]. In the presence of an additive (HNO3), the current Fe(III) catalytic system shows significant increase in the total yield, whereas the V(V) systems are effective under additive-free conditions [27,42]. The total yield under US conditions is also higher than that shown by a hydrazone Cu(II) catalytic system under MW conditions [27]. The selectivity towards CyOH of the current catalytic system is moderate in comparison with the hydrazone V(V) system [35,42].
From radical trapping experiments (addition of 2,2,6,6-tetramethylpiperidin-l-oxyl free radical to the reaction medium), in the presence of 1, 2, or 3 (Table 2, entries 21, 24, and 27, respectively), an extensive product yield inhibition (over 80%), relative to the total yield obtained when the reactions were carried out without any additive (Table 2, entries 5, 22, and 25, respectively), was observed. This suggests that the cyclohexane oxidation catalysed by complexes 1-3 proceeds through a radical mechanism, as proposed in other cases, depicted in Equations (1)–(9) [27,35,36,54,62]. Firstly, the reaction proceeds through the iron-catalyzed decomposition of the oxidant, leading to the formation of t-BuOO and t-BuO radicals upon the reduction of Fe(III) and oxidation of Fe(II) species, according to reactions (1) and (2), respectively. Then, cyclohexyl radical (Cy) formation takes place due to H-abstraction from CyH by t-BuO (reaction (3)). Upon the reaction with dioxygen, Cy forms CyOO (reaction (4)), and then upon H-abstraction from TBHP by CyOO, CyOOH is formed (reaction (5)). In the reactions (6) and (7), the Fe-assisted decomposition of CyOOH produces CyO and CyOO, which leads to the formation of cyclohexanol (CyOH) and cyclohexanone (Cy=O) (the desired products), according to reactions (8) and (9).
[FeIII]+ t-BuOOH → t-BuOO + H+ + [FeII]
[FeII]+ t-BuOOH → t-BuO + [FeIII] + HO
t-BuO + CyH → t-BuOH + Cy
Cy + O2 → CyOO
CyOO + t-BuOOH → CyOOH + t-BuOO
CyOOH + [FeII] → CyO + HO + [FeIII]
CyOOH + [FeIII]→ CyOO + H+ + [FeII]
CyO + CyH→ CyOH + Cy
2CyOO → CyOH + Cy=O + O2

3. Experimental

3.1. General Materials and Procedures

The syntheses were performed in air. Reagents and solvents were obtained from commercial sources and used as received, without further purification or drying. Fe(NO3)3.9H2O was used as the source of metal ion for the synthesis of 1. Anhydrous FeCl2 was used for the syntheses of 2 and 3.
C, H, and N elemental analyses were carried out by the Microanalytical Service of the Instituto Superior Técnico. Infrared spectra (4000–400 cm−1) were recorded on a Bruker Vertex 70 instrument (Bruker Corporation, Ettlingen, Germany) in KBr pellets; wavenumbers are given in cm−1. Mass spectra were run in a Varian 500-MS LC Ion Trap Mass Spectrometer (Agilent technologies, Amstelveen, The Netherlands) equipped with an electrospray (ESI) ion source. The drying gas and flow rate were optimized (for electrospray ionization) according to the particular sample, with a 35 p.s.i. nebulizer pressure. Scanning was performed from m/z 100 to 1200 in methanol solution. The compounds were observed in the positive mode (capillary voltage = 80–105 V).

3.2. Synthesis of the Pro-Ligand H2L

-acetylpyrazine-2-carbohydrazide (H2L), the pro-ligand, (Scheme 1), was synthesized as described in the literature [45], by the acetylation of pyrazine-2-carbohydrazide.
Yield: 86.0%. Anal. calc. for (C7H8N4O2): C, 46.67; H, 4.48; N, 31.10; found: C, 46.62; H, 4.43; N, 31.06%. IR (KBr pellet, cm−1): 3324 ν(NH), 3216 ν(NH), 1698 ν(C=O), 1670 ν(C=O). 1H NMR (DMSO-d6, δ): 9.18-8.86 (m, 3H, C4H3N2), 8.74 (s, 2H, NH), 1.91 (s, 3H, CH3).

3.3. Syntheses of Fe(III) Complexes of Nʹ-acetylpyrazine-2-carbohydrazide

3.3.1. [Fe(HL)(H2O)2(NO3)]NO3 (1)

In this study, 0.404 g (1.00 mmol) of Fe(NO3)3.9H2O was added to a 30 mL methanol solution of H2L (0.180 g, 1.00 mmol). Then, at room temperature, the reaction mixture was stirred for 30 min, in open air. The resultant reddish brown solution was filtered and the filtrate was kept in air. Single crystals of 1 were isolated after 3 days, washed twice with cold CHCl3, and dried in open air.
Yield 66%. Anal. Calcd. for C7H11FeN6O10: C, 21.28; H, 2.81; N, 21.27. Found: C, 21.20; H, 2.72; N, 21.18. IR (KBr pellet, cm−1): 3136 ν(NH), 1383 ν(NO3)-, 1678 ν(C=O), 1646 ν(C=O), 1032 ν(N–N). ESI-MS (+): m/z 333 [M-(NO3)]+ (100%).

3.3.2. [Fe(HL)Cl2] (2) and [Fe(HL)Cl(µ-OMe)]2 (3)

Both 2 and 3 were synthesized according to the literature [45].
[Fe(HL)Cl2] (2)
Yield 70%. Anal. Calcd. for C7H7Cl2FeN4O2: C, 27.48; H, 2.31; N, 18.31. Found: C, 27.42; H, 2.28; N, 18.25. IR (KBr pellet, cm−1): 3134 ν(NH), 1702 ν(C=O), 1668 ν(C=O), 1037 ν(N–N). ESI-MS (+): m/z 306 [M+H]+ (100%).
[Fe(HL)Cl(µ-OMe)]2 (3)
Yield 65%. Anal. Calcd. for C16H20Cl2Fe2N8O6: C, 31.87; H, 3.34; N, 18.58. Found: C, 31.84; H, 3.33; N, 18.54. IR (KBr; cm−1): 3138 ν(NH), 1664 ν(C=O), 1641 ν(C=O), 1038 ν(N–N). ESI-MS (+): m/z 604 [M+H]+ (100%).

3.4. X-ray Measurements

A crystal of 1 was immersed in cryo-oil, mounted in a Nylon loop, and measured at 150 K. Intensity data were collected using a Bruker AXS-KAPPA APEX II diffractometer (Bruker AXS Inc., Madison, WI, USA) with graphite monochromated Mo-Kα (λ 0.71073) radiation. Omega scans of 0.5° per frame were used for data collection and a full sphere of data was obtained. Cell parameters were retrieved using Bruker SMART [63] software and the data were refined using Bruker SAINT [63] on all the observed reflections. SADABS was used for absorption corrections [64]. Direct methods were employed by using SHELXS97 [65] and refined with SHELXL2018/3 [66]. Calculations were conducted using WinGX version 2018.3 [67]. The H-atoms bonded to carbon and nitrogen were introduced in the model at geometrically calculated positions and refined using a riding model, with Uiso defined as 1.2Ueq of the parent carbon atoms for aromatic residues and for nitrogen, and 1.5Ueq for the methyl group. The hydrogen atoms of water ligands were found in the difference Fourier map and allowed to refine with distance restrains. The non-coordinated water molecule was disordered over two positions and refined with the use of PART instruction, with the occupancy of O1W and O2W at a ratio of 74% and 26%, respectively. After this strategy, the remaining electron density was considered as being due to another water molecule (O3W), which was refined with an occupancy of 0.25, flanked by PART -1, and the structure finalized normally. The hydrogen atoms of O1W, O2W, and O3W could be neither located nor inserted in the calculated positions (the use of the CALC-OH routine in WinGX proved to be unsuccessful). Least square refinements were carried out with anisotropic thermal motion parameters for all of the non-hydrogen atoms.

3.5. Catalytic Studies

The microwave-assisted peroxidative oxidation of cyclohexane was carried out in G10 Pyrex tubes (10 mL capacity reaction tube with a 13 mm internal diameter) in a focused Anton Paar Monowave 300 reactor (Anton Paar GmbH, Graz, Austria) fitted with a rotational system and an IR temperature detector.
The test was performed under the following conditions: The desired amount of catalyst 1, 2, or 3 (2.5–20 μmol); cyclohexane (5 mmol); acetonitrile (3 mL) and oxidant (1, 2, or 4 equivalent vs. substrate, t-BuOOH, 70% aqueous solution) were added into an Pyrex tube, which was introduced in the MW reactor and stirred under irradiation (5-–10 W) at 50 °C. After completion of the desired reaction time, the reaction mixture was cooled to room temperature, whereafter 90 µL of cycloheptanone (internal standard) and 10 mL of diethyl ether (for substrate and organic product extraction) were added. The reaction mixture was stirred and centrifuged. A sample was taken from the mixture and analyzed by gas chromatography (GC) after the addition of an excess of triphenylphosphine, in order to reduce cyclohexyl hydroperoxide to cyclohexanol, following a method developed by Shul’pin [54]. The composition of products was confirmed by Gas Chromatography-Mass Spectrometry (GC-MS).
For the ultrasound-assisted oxidation reactions, a reaction tube was immersed in an ATU ultrasonic thermoregulated bath (40 kHz, 600 W) equipped with an automatic temperature heating-cooling circulatory system, which kept the bath temperature at ca. 50 °C for all of the trials.
GC measurements were carried out using a FISONS Instruments GC 8000 series gas chromatograph (Agilent Technologies, Santa Clara, CA, USA) with an FID detector and a capillary column (DB-WAX, column length: 30 m; internal diameter: 0.32 mm), using helium as a carrier gas and the Jasco-Borwin v.1.50 software (Jasco, Tokyo, Japan). The samples were injected at 240 °C, whereas the initial temperature was maintained at 100 °C for 1 min, increased to 180 °C at the rate 10 °C/min, and then hold for 1 min.
The samples were analysed by GC-MS using a Perkin Elmer Clarus 600 C instrument (Shelton, CT, USA) (He as the carrier gas). The ionization voltage was 70 eV. Gas chromatography was guided in the temperature-programming mode, using an SGE BPX5 column (30 m × 0.25 mm × 0.25 µm).
Retention times of all identified reaction products were compared with those of the commercially available samples. Reaction products’ mass spectra were compared to fragmentation patterns obtained from the NIST spectral library stored in the computer software of the mass spectrometer.

4. Conclusions

Three Fe(III) compounds (1-3) derived from -acetylpyrazine-2-carbohydrazide were successfully applied for the peroxidative oxidation of cyclohexane under different energy stimuli (microwave irradiation, ultrasound, and conventional heating). In order to improve the catalytic performance, the effects of different reaction parameters were studied, namely the reaction time, catalyst amount, substrate:oxidant ratio, and presence of additives. Compounds 1-3 catalyzed the oxidation of cyclohexane via a radical mechanism, yielding cyclohexanol and cyclohexanone as the main products; the former with a high selectivity upon the reduction of the primary product cyclohexylperoxide (up to 95%). Complexes 1 and 3 exhibited similar catalytic activities, with 14–15% yields of cyclohexanol and cyclohexanone under 3 h of microwave irradiation at 50 °C, whereas complex 2 reached a total product yield of 10.4%.
Apart from being undertaken under microwaves, the peroxidative oxidation of cyclohexane was also performed using other types of energy inputs, i.e., conventional heating and ultrasound irradiation. The catalysts responded differently to the various energy stimuli, with the best performance of 1 being observed under microwave radiation, whereas complex 2 showed the maximum activity under conventional heating and compound 3 under ultrasounds. However, in all cases, the highest selectivity to cyclohexanol was verified when the reaction was assisted by microwave radiation.
In the case of 1, a significant increase of the total yield was observed from 15 to 38% in the presence of an additive (HNO3), although such an influence was not found for 2 and 3. Therefore, the combined use of nitric acid and the mononuclear cationic Fe(III) compound 1 plays a crucial role in accelerating the catalytic oxidation process.
This study can help to foster the fruitful use of the environmentally acceptable oxidant aqueous TBHP and the application of microwave heating or ultrasounds to promote catalysis, which have significant environmental implications.

Supplementary Materials

CCDC number 2019464 contains the supplementary crystallographic data for 1. This can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html, or from the Cambridge Crystallographic Data Center, 12 Union Road, Cambridge CB2 1EZ, UK; fax: (+44)-1223-336-033; or e-mail: [email protected].

Author Contributions

Conceptualization, M.S. and E.C.B.A.A.; Methodology, M.S. and E.C.B.A.A.; Software, M.S., M.d.F.C.G.d.S. and T.R.B.; Validation, M.S., M.d.F.C.G.d.S. and E.C.B.A.A.; Formal Analysis, M.S. and T.R.B.; Investigation, M.S. and T.R.B; Resources, M.S. and A.J.L.P.; Data Curation, M.S., M.d.F.C.G.d.S. and T.R.B.; Writing-Original Draft Preparation, M.S., T.R.B. and E.C.B.A.A.; Writing-Review & Editing, M.d.F.C.G.d.S. and A.J.L.P.; Visualization, M.S.; Supervision, M.S. and E.C.B.A.A.; Project Administration, A.J.L.P.; Funding Acquisition, A.J.L.P and E.C.B.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been supported by the Fundação para a Ciência e Tecnologia (FCT) 2020-2023 multiannual funding to Centro de Química Estrutural (project UIDB/00100/2020).

Acknowledgments

The authors are grateful to the Fundação para a Ciência e Tecnologia (FCT) project PTDC/QUI-QIN/29778/2017 for financial support. M.S. acknowledges the FCT and IST for a working contract "DL/57/2017" (Contract no. IST-ID/102/2018).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pombeiro, A.J.L.; Guedes da Silva, M.F.C. (Eds.) Alkane Functionalization; Wiley: Hoboken, NJ, USA, 2019. [Google Scholar]
  2. Pombeiro, A.J.L. Alkane Functionalization: Introduction and overview. In Alkane Functionalization; Pombeiro, A.J.L., Guedes da Silva, M.F.C., Eds.; Wiley: Hoboken, NJ, USA, 2019; Chapter 1; pp. 1–15. [Google Scholar]
  3. Shilov, A.E.; Shul’pin, G.B. Activation of C-H bonds by metal complexes. Chem. Rev. 1997, 97, 2879–2932. [Google Scholar] [CrossRef] [PubMed]
  4. Shul’pin, G.B. Selectivity enhancement in functionalization of C-H bonds: A review. Org. Biomol. Chem. 2010, 8, 4217–4228. [Google Scholar] [CrossRef] [PubMed]
  5. Sutradhar, M.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Vanadium complexes: Recent progress in oxidation catalysis. Coord. Chem. Rev. 2015, 301, 200–239. [Google Scholar] [CrossRef]
  6. Shul’pin, G.B. Transition Metals for Organic Synthesis, 2nd ed.; Beller, M., Bolm, C., Eds.; Wiley-VCH: New York, NY, USA, 2004; Volume 2, Chapter 2; pp. 215–242. [Google Scholar]
  7. Shilov, A.E.; Shul’pin, G.B. Activation and Catalytic Reactions of Saturated Hydrocarbons in the Presence of Metal Complexes; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2000. [Google Scholar]
  8. 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]
  9. Derouane, E.G.; Haber, J.; Lemos, F.; Ribeiro, F.R.; Guinet, M. (Eds.) Catalytic Activation and Functionalization of Light Alkanes; NATO ASI Series; Kluwer Academic Publisher: Dordrecht, The Netherlands, 1998; Volume 44. [Google Scholar]
  10. Retcher, B.; Sánchez Costa, J.; Tang, J.; Hage, R.; Gamez, P.; Reedijk, J. Unexpected high oxidation of cyclohexane by Fe salts and dihydrogen peroxide in acetonitrile. J. Mol. Catal. A Chem. 2008, 286, 1–5. [Google Scholar] [CrossRef]
  11. Antony, R.; Manickam, T.S.; Kollu, P.; Chandrasekar, P.V.; Karuppasamy, K.; Balakumar, S. Highly dispersed Cu(II), Co(II) and Ni(II) catalysts covalently immobilized on imine-modified silica for cyclohexane oxidation with hydrogen peroxide. RSC Adv. 2014, 4, 24820–24830. [Google Scholar] [CrossRef]
  12. Rahman, A.; Mupa, M.; Mahamadi, C. A mini review on new emerging trends for the synthesis of adipic acid from metal-nano heterogeneous catalysts. Catal. Lett. 2016, 146, 788–799. [Google Scholar] [CrossRef]
  13. Guo, X.; Xu, M.; She, M.; Zhu, Y.; Shi, T.; Chen, Z.; Peng, L.; Guo, X.; Lin, M.; Ding, W. Morphology-reserved synthesis of discrete nanosheets of CuO@SAPO-34 and pore mouth catalysis for one-pot oxidation of cyclohexane. Angew. Chem. Int. Ed. Engl. 2020, 59, 2606–2611. [Google Scholar] [CrossRef]
  14. Schuchardt, U.; Cardoso, D.; Sercheli, R.; Pereira, R.; Cruz, R.S.; Guerreiro, M.C.; Pires, E.L. Cyclohexane oxidation continues to be a challenge. Appl. Catal. A Gen. 2001, 211, 1–17. [Google Scholar] [CrossRef]
  15. Pokutsa, A.; Le Bras, J. Muzart, Glyoxal-promoted homogeneous catalytic oxygenation of cyclohexane with hydrogen peroxide in the presence of V and Co compounds. J. Russ. Chem. Bull. Int. Ed. 2005, 54, 312–315. [Google Scholar] [CrossRef]
  16. Pombeiro, A.J.L. (Ed.) Advances in Organometallic Chemistry and Catalysis, The Silver/Gold Jubilee ICOMC Celebratory Book; J.Wiley & Sons: New York, NY, USA, 2014. [Google Scholar]
  17. Shul’pin, G.B. Hydrocarbon oxygenations with peroxides catalyzed by metal compounds. Mini Rev. Org. Chem. 2009, 6, 95–104. [Google Scholar] [CrossRef]
  18. Li, J.J. (Ed.) C-H Bond Activation in Organic Synthesis; CRC Press: Boca Raton, FL, USA, 2015. [Google Scholar]
  19. Pérez, P.J. (Ed.) Alkane C-H Activation by Single-Site Metal Catalysis; Springer: Berlin, Germany, 2012. [Google Scholar]
  20. Bäckvall, J.-E. Modern Oxidation Methods; Wiley: Hoboken, NJ, USA, 2011. [Google Scholar]
  21. White, M.C. Chemistry. Adding aliphatic C-H bond oxidations to synthesis. Science 2012, 335, 807–809. [Google Scholar] [CrossRef] [PubMed]
  22. Newhouse, T.; Baran, P.S. If C-H bonds could talk: Selective C-H bond oxidation. Angew. Chem. Int. Ed. Engl. 2011, 123, 3422–3435. [Google Scholar] [CrossRef]
  23. Olah, G.A.; Molnar, A.; Surya Prakash, G.K. Hydrocarbon Chemistry, 3rd ed.; Wiley: Hoboken, NJ, USA, 2017. [Google Scholar]
  24. Sutradhar, M.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Liu, C.-M.; Pombeiro, A.J.L. Trinuclear Cu(II) structural isomers: Coordination, magnetism, electrochemistry and catalytic activity toward oxidation of alkanes. Eur. J. Inorg. Chem. 2015, 2015, 3959–3969. [Google Scholar] [CrossRef]
  25. Sutradhar, M.; Roy Barman, T.; Alegria, E.C.B.A.; Guedes da Silva, M.F.C.; Liu, C.-M.; Kou, H.-Z.; Pombeiro, A.J.L. Cu(II) complexes of N-rich aroylhydrazone: Magnetism and catalytic activity towards microwave-assisted oxidation of xylenes. Dalton Trans. 2019, 48, 12839–12849. [Google Scholar] [CrossRef] [PubMed]
  26. Sutradhar, M.; Alegria, E.C.B.A.; Guedes da Silva, M.F.C.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Aroylhydrazone Cu(II) complexes in keto form: Structural characterization and catalytic activity towards cyclohexane oxidation. Molecules 2016, 21, 425. [Google Scholar] [CrossRef] [Green Version]
  27. Sutradhar, M.; Alegria, E.C.B.A.; Guedes da Silva, M.F.C.; Liu, C.-M.; Pombeiro, A.J.L. Peroxidative oxidation of alkanes and alcohols under mild conditions by di- and tetranuclear copper(II) complexes of bis(2-hydroxybenzylidene)isophthalohydrazide. Molecules 2018, 23, 2699. [Google Scholar] [CrossRef] [Green Version]
  28. Bonchio, M.; Carraro, M.; Scorrano, G.; Kortz, U. Microwave-assisted fast cyclohexane oxygenation catalyzed by iron-substituted polyoxotungstates. Adv. Synth. Catal. 2005, 347, 1909–1912. [Google Scholar] [CrossRef]
  29. Carvalho, N.M.; Alvarez, H.M.; Horn, A., Jr.; Antunes, O.A. Influence of microwave irradiation in the cyclohexane oxidation catalyzed by Fe(III) complexes. Catal. Today 2008, 133, 689–694. [Google Scholar] [CrossRef]
  30. Fernandes, R.; Lasri, J.; Guedes da Silva, M.F.C.; da Silva, J.A.L.; Pombeiro, A.J.L. Bis- and tris-pyridyl amino and imino thioether Cu and Fe complexes. Thermal and microwave-assisted peroxidative oxidations of 1-phenylethanol and cyclohexane in the presence of various N-based additives. J. Mol. Catal. A Chem. 2011, 351, 100–111. [Google Scholar] [CrossRef]
  31. 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]
  32. Ribeiro, A.P.C.; Martins, L.M.D.R.S.; Carabineiro, S.A.C.; Buijnsters, J.G.; Figueiredo, J.L.; Pombeiro, A.J.L. Heterogenised C-scorpionate iron(II) complex on nanostructured carbon materials as catalysts for microwave-assisted oxidation reactions. ChemCatChem 2018, 10, 1821–1828. [Google Scholar] [CrossRef]
  33. Ribeiro, A.P.C.; Matias, I.A.S.; Alegria, E.C.B.A.; Ferraria, A.M.; Botelho do Rego, A.M.; Pombeiro, A.J.L.; Martins, L.M.D.R.S. New trendy magnetic C-scorpionate iron catalyst and its performance towards cyclohexane oxidation. Catalysts 2018, 8, 69. [Google Scholar] [CrossRef] [Green Version]
  34. Sutradhar, M.; Kirillova, M.V.; Guedes da Silva, M.F.C.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. A hexanuclear mixed-valence oxovanadium(IV,V) complex as a highly efficient alkane oxidation catalyst. Inorg. Chem. 2012, 51, 11229–11231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Sutradhar, M.; Martins, L.M.; Roy, B.T.; Kuznetsov, M.L.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Vanadium complexes of different nuclearities in the catalytic oxidation of cyclohexane and cyclohexanol—An experimental and theoretical investigation. New. J. Chem. 2019, 43, 17557–17570. [Google Scholar] [CrossRef]
  36. Sutradhar, M.; Martins, L.M.D.R.S.; Carabineiro, S.A.C.; Guedes da Silva, M.F.C.; Buijnsters, J.G.; Figueiredo, J.L.; Pombeiro, A.J.L. Oxidovanadium(V) complexes anchored on carbon materials as catalysts for the oxidation of 1-phenylethanol. ChemCatChem 2016, 8, 2254–2266. [Google Scholar] [CrossRef]
  37. Sutradhar, M.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Alegria, E.C.B.A.; Liu, C.-M.; Pombeiro, A.J.L. Mn(II,II) complexes: Magnetic properties and microwave assisted oxidation of alcohols. Dalton Trans. 2014, 43, 3966–3977. [Google Scholar] [CrossRef]
  38. Sutradhar, M.; Roy Barman, T.; Pombeiro, A.J.L.; Martins, L.M.D.R.S. Catalytic activity of polynuclear vs. dinuclear aroylhydrazone Cu(II) complexes in microwave-assisted oxidation of neat aliphatic and aromatic hydrocarbons. Molecules 2019, 24, 47. [Google Scholar] [CrossRef] [Green Version]
  39. Sutradhar, M.; Alegria, E.C.B.A.; Mahmudov, K.T.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Iron(III) and cobalt(III) complexes with both tautomeric (keto and enol) forms of aroylhydrazone ligands: Catalysts for the microwave assisted oxidation of alcohols. RSC Adv. 2016, 6, 8079–8088. [Google Scholar] [CrossRef]
  40. Zaltariov, M.-F.; Alexandru, M.; Cazacu, M.; Shova, S.; Novitchi, G.; Train, C.; Dobrov, A.; Kirillova, M.V.; Alegria, E.C.B.A.; Pombeiro, A.J.L.; et al. Tetranuclear copper(II) complexes with macrocyclic and open-chain disiloxane ligands as catalyst precursors for hydrocarboxylation and oxidation of alkanes and 1-phenylethanol. Eur. J. Inorg. Chem. 2014, 29, 4946–4956. [Google Scholar] [CrossRef]
  41. Dobrov, A.; Darvasiová, D.; Zalibera, M.; Bučinský, L.; Puškárová, I.; Rapta, P.; Martins, L.M.D.R.S.; Pombeiro, A.J.L.; Arion, V.B. Nickel(II) complexes with redox noninnocent octaazamacrocycles as catalysts in oxidation reactions. Inorg. Chem. 2019, 58, 11133–11145. [Google Scholar] [CrossRef]
  42. 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] [PubMed]
  43. Arion, V.B.; Platzer, S.; Rapta, P.; Machata, P.; Breza, M.; Vegh, D.; Dunsch, L.; Telser, J.; Shova, S.; Mac Leod, T.; et al. Marked stabilization of redox states and enhanced catalytic activity in galactose oxidase models based on transition metal S-methylisothiosemicarbazonates with—SR group in ortho-position to the phenolic oxygen. Inorg. Chem. 2013, 52, 7524–7540. [Google Scholar] [CrossRef] [PubMed]
  44. Dobrov, A.; Fesenko, A.; Yankov, A.; Stepanenko, I.; Darvasiová, D.; Breza, M.; Rapta, P.; Martins, L.M.D.R.S.; Pombeiro, A.J.L.; Shutalev, A.; et al. Nickel(II), Copper(II) and Palladium(II) complexes with Bis-Semicarbazide hexaazamacrocycles: Redox-noninnocent behavior and catalytic activity in oxidation and C-C coupling reactions. Inorg. Chem. 2020, 59, 10650–10664. [Google Scholar] [CrossRef] [PubMed]
  45. Roy Barman, T.; Sutradhar, M.; Alegria, E.C.B.A.; Guedes da Silva, M.F.C.; Kuznetsov, M.L.; Pombeiro, A.J.L. Efficient Solvent-Free Friedel-Crafts Benzoylation and Acylation of m-Xylene Catalyzed by N-acetylpyrazine-2-carb hydrazide-Fe(III)-chloro Complexes. Chem. Select 2018, 3, 8349–8355. [Google Scholar]
  46. Dudley, G.B.; Richert, R.; Stiegman, A.E. On the existence of and mechanism for microwave-specific reaction rate enhancement. Chem. Sci. 2015, 6, 2144–2152. [Google Scholar] [CrossRef]
  47. Varma, R.S. Journey on greener pathways: From the use of alternate energy inputs and benign reaction media to sustainable applications of nano-catalysts in synthesis and environmental remediation. Green Chem. 2014, 16, 2027–2041. [Google Scholar] [CrossRef]
  48. Ribeiro, A.P.C.; Alegria, E.C.B.A.; Palavra, A.; Pombeiro, A.J.L. Alkane functionalization under unconventional conditions: In ionic liquid, in supercritical CO2 and microwave assisted. In Alkane Functionalization; Pombeiro, A.J.L., Guedes da Silva, M.F.C., Eds.; Wiley: Hoboken, NJ, USA, 2019; Chapter 24; pp. 523–537. [Google Scholar]
  49. Ribeiro, A.P.C.; Alegria, E.C.B.A.; Kopylovich, M.N.; Ferraria, A.M.; Botelho do Rego, A.M.; Pombeiro, A.J.L. On the comparison of microwave and mechanochemical energy inputs in catalytic oxidation of cyclohexane. Dalton Trans. 2018, 47, 8193–8198. [Google Scholar] [CrossRef]
  50. Perkas, N.; Wang, Y.; Koltypin, Y.; Gedanken, A.; Chandrasekaran, S. Mesoporous iron–titania catalyst for cyclohexane oxidation. Chem. Commun. 2001, 988–989. [Google Scholar] [CrossRef]
  51. Sutradhar, M.; Roy Barman, T.; Pombeiro, A.J.L.; Martins, L.M.D.R.S. Cu(II) and Fe(III) complexes derived from N-acetylpyrazine-2-carbohydrazide as efficient catalysts towards solvent-free microwave assisted oxidation of alcohols. Catalysts 2019, 9, 1053. [Google Scholar] [CrossRef] [Green Version]
  52. Sutradhar, M.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Synthesis and chemical reactivity of an Fe(III) metallacrown-6 towards N-donor Lewis bases. Inorg. Chem. Commun. 2013, 30, 42–45. [Google Scholar] [CrossRef]
  53. Sutradhar, M.; Guedes da Silva, M.F.C.; Nesterov, D.S.; Jezierska, J.; Pombeiro, A.J.L. 1D coordination polymer with octahedral and square-planar nickel(II) centers. Inorg. Chem. Commun. 2013, 29, 82–84. [Google Scholar] [CrossRef]
  54. Shul’pin, G.B.; Nizova, G.V. Formation of alkyl peroxides in oxidation of alkanes by H2O2 catalyzed by transition metal complexes React. Kinet. Catal. Lett. 1992, 48, 333–338. [Google Scholar] [CrossRef]
  55. Pokutsa, A.; Pawel Bloniarz, P.; Fliunt, O.; Kubaj, Y.; Zaborovskyia, A.; Paczeŝniakc, T. Sustainable oxidation of cyclohexane catalyzed by a VO(acac)2-oxalic acid tandem: The electrochemical motive of the process efficiency. RSC Adv. 2020, 10, 10959–10971. [Google Scholar] [CrossRef] [Green Version]
  56. Shul’pin, G.S.; Mishra, L.S.; Shul’pina, T.V.; Strelkova, A.J.L. Pombeiro. Oxidation of hydrocarbons with hydrogen peroxide catalysed by maltolato vanadium complexes covalently bonded to silica gel. Catal. Commun. 2007, 8, 1516–1520. [Google Scholar] [CrossRef]
  57. Sutradhar, M.; Alegria, E.C.B.A.; Barman, T.R.; Scorcelletti, F.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Microwave-assisted peroxidative oxidation of toluene and 1-phenylethanol with monomeric keto and polymeric enol aroylhydrazone Cu(II) complexes. Mol. Catal. 2017, 439, 224–232. [Google Scholar] [CrossRef]
  58. Días-Ortiz, Á.; Prieto, P.; de la Hoz, A. A critical overview on the effect of microwave irradiation in organic synthesis. Chem. Rec. 2019, 19, 85–97. [Google Scholar] [CrossRef] [Green Version]
  59. Herrero, M.A.; Kremsner, J.M.; Kappe, C.O. Nonthermal microwave effects revisited: On the importance of internal temperature monitoring and agitation in microwave chemistry. J. Organomet. Chem. 2007, 73, 36–47. [Google Scholar] [CrossRef]
  60. Obermayer, D.; Kappe, C.O. On the importance of simultaneous infrared/fiber-optic temperature monitoring in the microwave-assisted synthesis of ionic liquids. Org. Biomol. Chem. 2010, 8, 114–121. [Google Scholar] [CrossRef]
  61. Piermattei, A.; Karthikeyan, S.; Sijbesma, R.P. Activating catalysts with mechanical force. Nat. Chem. 2009, 1, 133–137. [Google Scholar] [CrossRef]
  62. Shul’pin, G.B. Metal-catalysed hydrocarbon oxidations. C. R. Chim. 2003, 6, 163–178. [Google Scholar] [CrossRef]
  63. Bruker AXS Inc. Bruker, APEX2; Bruker AXS Inc.: Madison, Wisconsin, USA, 2012. [Google Scholar]
  64. Sheldrick, G.M. SADABS. In Program for Empirical Absorption Correction; University of Göttingen: Göttingen, Germany, 2000. [Google Scholar]
  65. Sheldrick, G.M. SHELX97. In Programs for Crystal Structure Analysis (Release 97-2); University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  66. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Farrugia, L.J. wingx and ortep for windows: An update. J. Appl. Cryst. 2012, 45, 849–854. [Google Scholar] [CrossRef]
Scheme 1. Syntheses of 13.
Scheme 1. Syntheses of 13.
Catalysts 10 01175 sch001
Figure 1. Ellipsoid plot of the complex cation in 1 (drawn at a 30% probability level) with an atom labeling scheme. Selected bond distances (Å) and angles (°): N3–N4 1.389(3); C5–O2 1.241(3); C6–O1 1.266(3); N1–Fe1 2.2547(19); N3–Fe1 2.034(2); O1–Fe1 2.0707(17); O3–Fe1 1.9998(17); O4–Fe1 1.9910(17); O5–Fe1 2.1163(17); O6–Fe1 2.2508(18); O4–Fe1–O3 167.97(8); O1–Fe1–N1 146.91(7); N3–Fe1–N1 72.68(7); N3–Fe1–O1 74.29(7); and O3–Fe1–N3 95.88(7).
Figure 1. Ellipsoid plot of the complex cation in 1 (drawn at a 30% probability level) with an atom labeling scheme. Selected bond distances (Å) and angles (°): N3–N4 1.389(3); C5–O2 1.241(3); C6–O1 1.266(3); N1–Fe1 2.2547(19); N3–Fe1 2.034(2); O1–Fe1 2.0707(17); O3–Fe1 1.9998(17); O4–Fe1 1.9910(17); O5–Fe1 2.1163(17); O6–Fe1 2.2508(18); O4–Fe1–O3 167.97(8); O1–Fe1–N1 146.91(7); N3–Fe1–N1 72.68(7); N3–Fe1–O1 74.29(7); and O3–Fe1–N3 95.88(7).
Catalysts 10 01175 g001
Scheme 2. Microwave (MW)-assisted oxidation of cyclohexane (CyH).
Scheme 2. Microwave (MW)-assisted oxidation of cyclohexane (CyH).
Catalysts 10 01175 sch002
Figure 2. Oxygenated product yield of cyclohexane (CyH) oxidation [cyclohexanol (CyOH) + cyclohexanone (Cy=O)] with respect to the following: (a) reaction time (10 μmol of 1); (b) catalyst amount (3 h reaction time); (c) acid amount (10 μmol of 1 and 3 h reaction time); and (d) reaction time in the presence of acid additive [10 μmol of 1 and n(HNO3)/nʹ(catalyst) = 25]. Other reaction conditions: CyH (5 mmol); NCMe (3 mL); TBHP (70% aqueous) (10 mmol); 50 °C; under MW-irradiation (5 W); and GC analysis after the addition of PPh3.
Figure 2. Oxygenated product yield of cyclohexane (CyH) oxidation [cyclohexanol (CyOH) + cyclohexanone (Cy=O)] with respect to the following: (a) reaction time (10 μmol of 1); (b) catalyst amount (3 h reaction time); (c) acid amount (10 μmol of 1 and 3 h reaction time); and (d) reaction time in the presence of acid additive [10 μmol of 1 and n(HNO3)/nʹ(catalyst) = 25]. Other reaction conditions: CyH (5 mmol); NCMe (3 mL); TBHP (70% aqueous) (10 mmol); 50 °C; under MW-irradiation (5 W); and GC analysis after the addition of PPh3.
Catalysts 10 01175 g002
Figure 3. Effects of different additives on the oxidation of cyclohexane catalyzed by 1. Reaction conditions: CyH (5 mmol); 1 (10 μmol); TBHP (70% aq.) (10 mmol); NCMe (3 mL); 50 °C; 3 h; under MW-irradiation (5 W).
Figure 3. Effects of different additives on the oxidation of cyclohexane catalyzed by 1. Reaction conditions: CyH (5 mmol); 1 (10 μmol); TBHP (70% aq.) (10 mmol); NCMe (3 mL); 50 °C; 3 h; under MW-irradiation (5 W).
Catalysts 10 01175 g003
Figure 4. Effect of different energy stimuli on the oxidation of cyclohexane catalyzed by 1. Reaction conditions: CyH (5 mmol); 1 (10 μmol); TBHP (70% aqueous) (10 mmol); NCMe (3 mL); n(HNO3)/nʹ(catalyst) = 25; 50 °C; 0.5-6 h. CONV = under conventional heating with oil bath; MW = under MW-assisted condition; US = under ultrasound irradiation.
Figure 4. Effect of different energy stimuli on the oxidation of cyclohexane catalyzed by 1. Reaction conditions: CyH (5 mmol); 1 (10 μmol); TBHP (70% aqueous) (10 mmol); NCMe (3 mL); n(HNO3)/nʹ(catalyst) = 25; 50 °C; 0.5-6 h. CONV = under conventional heating with oil bath; MW = under MW-assisted condition; US = under ultrasound irradiation.
Catalysts 10 01175 g004
Table 1. Crystallographic parameters and structure refinement data for complex 1.
Table 1. Crystallographic parameters and structure refinement data for complex 1.
1
Empirical formulaC7H11FeN6O11.25
Formula weight415.04
Crystal systemMonoclinic
Space groupP21/c
Temperature/K150(2)
a7.2541(4)
b14.4633(8)
c14.7415(9)
β/°93.275(2)
V3)1544.13(15)
Z4
Dcalc (g cm–3)1.785
F000844
μ(Mo Kα) (mm–1)1.052
Rfls. collected/unique/observed22957/2837/2385
Rint0.0478
Final R1a, wR2b (I ≥ 2σ)0.0319, 0.0711
Goodness-of-fit on F21.051
a R = Σ||Fo|–|Fc||/Σ|Fo|; b wR(F2) = [Σw(|Fo|2 − |Fc|2)2/Σw|Fo|4]½.
Table 2. Data a for the MW-assisted oxidation of cyclohexane by catalysts 1-3 using TBHP (70% aqueous) as an oxidant.
Table 2. Data a for the MW-assisted oxidation of cyclohexane by catalysts 1-3 using TBHP (70% aqueous) as an oxidant.
EntryCatalystCatalyst Amount (mol)Reaction Time (h)Yield (%) bSelectivity to Cyclohexanol (%) d
CyOHCy=OTOTAL c
11100.51.60.11.794
21016.70.77.491
3 e1010.70.040.7495
4 f1037.53.110.671
510313.21.714.986
610614.22.716.984
7 g1037.61.18.787
8 h10316.74.821.578
92.533.60.23.895
10538.40.99.390
1120315.73.719.480
12 i10317.14.321.480
13 j10322.66.128.779
14 k10324.812.937.766
15 l10327.59.737.274
16 k1019.91.711.685
17 k10624.712.937.666
18 k10923.513.637.163
19 m1037.71.18.888
20 n10321.86.528.377
21 o1031.30.41.776
2221039.11.310.488
23 k1039.80.910.792
24 o1032.50.22.793
25310312.32.114.485
26 k10316.31.918.290
27 o1031.90.42.383
28Fe(NO3)3.9H2O1033.81.55.372
29Anhy. FeCl21033.30.94.279
a Reaction conditions (unless stated otherwise): cyclohexane (5.0 mmol); 2.5-20 µmol of catalyst; acetonitrile (3 mL); TBHP 70% aqueous solution (10 mmol); 1-9 h; 50 °C; microwave irradiation (5 W); yield and Turnover number (TON) determined by gas chromatography upon treatment with PPh3 (see text). b Molar yield (%) based on substrate, i.e., moles of product [cyclohexanol (CyOH) or cyclohexanone (Cy=O)] per 100 moles of cyclohexane after PPh3 treatment. c Total yield = moles of products [cyclohexanol (CyOH) + cyclohexanone (Cy=O)]/100 moles of cyclohexane. d Selectivity to cyclohexanol relative to KA oil mixture, i.e., moles of CyOH/(100 moles of CyOH + Cy=O). e Solvent-free conditions. f Without PPh3 treatment. g Oxidant:substrate = 1:1. h Oxidant:substrate = 4:1. i n(HNO3)/n(catalyst) = 5. j n(HNO3)/n(catalyst) = 10. k n(HNO3)/n(catalyst) = 25. l n(HNO3)/n(catalyst) = 50. m n(HPCA)/n(catalyst) = 25. n n(TFA)/n(catalyst) = 25. o n(TEMPO)/n(catalyst) = 25.
Table 3. Effect of different energy inputs for the oxidation of cyclohexane by catalysts 1-3 using TBHP (70% aq.) as an oxidant a.
Table 3. Effect of different energy inputs for the oxidation of cyclohexane by catalysts 1-3 using TBHP (70% aq.) as an oxidant a.
EntryCatalystMethodReaction Time (h)Yield (%) bSelectivity to Cyclohexanol (%) d
CyOHCy=OTOTAL c
11CONV0.516.55.121.676
2116.96.223.173
3318.98.627.569
4619.87.827.672
52428.811.840.671
6MW0.511.62.914.582
7 e115.24.619.877
8 e324.812.937.766
9 e624.712.937.666
10US0.516.44.921.377
11116.26.522.771
12315.47.723.167
13614.99.724.661
142CONV315.26.922.169
15MW39.80.910.792
16US37.34.011.365
173CONV313.34.417.775
18MW316.31.918.290
19US320.99.129.970
a Reaction conditions (unless stated otherwise): cyclohexane (5.0 mmol); 10 µmol of catalyst; acetonitrile (3 mL); TBHP 70% aqueous solution (10 mmol); n(HNO3)/nʹ(catalyst) = 25; 0.5-6 h; 50 °C; microwave irradiation (5 W); yield and TON determined by gas chromatography upon treatment with PPh3. b Molar yield (%) based on substrate, i.e., moles of product [cyclohexanol (CyOH) or cyclohexanone (Cy=O)] per 100 moles of cyclohexane after PPh3 treatment. c Total yield = moles of products [cyclohexanol (CyOH) + cyclohexanone (Cy=O)]/100 moles of cyclohexane. d Selectivity to cyclohexanol relative to KA oil mixture, i.e., moles of CyOH/(100 moles of CyOH + Cy=O). e Data from Table 2 (entries 14, 16, and 17). CONV = under conventional heating with oil bath; MW = under MW-assisted condition; US = under ultrasound irradiation.

Share and Cite

MDPI and ACS Style

Roy Barman, T.; Sutradhar, M.; C. B. A. Alegria, E.; C. Guedes da Silva, M.d.F.; Pombeiro, A.J.L. Fe(III) Complexes in Cyclohexane Oxidation: Comparison of Catalytic Activities under Different Energy Stimuli. Catalysts 2020, 10, 1175. https://doi.org/10.3390/catal10101175

AMA Style

Roy Barman T, Sutradhar M, C. B. A. Alegria E, C. Guedes da Silva MdF, Pombeiro AJL. Fe(III) Complexes in Cyclohexane Oxidation: Comparison of Catalytic Activities under Different Energy Stimuli. Catalysts. 2020; 10(10):1175. https://doi.org/10.3390/catal10101175

Chicago/Turabian Style

Roy Barman, Tannistha, Manas Sutradhar, Elisabete C. B. A. Alegria, Maria de Fátima C. Guedes da Silva, and Armando J. L. Pombeiro. 2020. "Fe(III) Complexes in Cyclohexane Oxidation: Comparison of Catalytic Activities under Different Energy Stimuli" Catalysts 10, no. 10: 1175. https://doi.org/10.3390/catal10101175

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