Next Article in Journal
Morphology- and Crystalline Composition-Governed Activity of Titania-Based Photocatalysts: Overview and Perspective
Next Article in Special Issue
Special Issue: Coordination Chemistry and Catalysis
Previous Article in Journal
Synthesis of Hydroxy Sodalite from Coal Fly Ash for Biodiesel Production from Waste-Derived Maggot Oil
Previous Article in Special Issue
The Catalytic Activities of Carbocyclic Fused Pyridineimine Nickel Complexes Analogues in Ethylene Polymerization by Modeling Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cu(II) and Fe(III) Complexes Derived from N-Acetylpyrazine-2-Carbohydrazide as Efficient Catalysts Towards Neat Microwave Assisted Oxidation of Alcohols

by
Manas Sutradhar
*,
Tannistha Roy Barman
,
Armando J. L. Pombeiro
and
Luísa M. D. R. S. Martins
*
Centro de Química Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisboa, Portugal
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(12), 1053; https://doi.org/10.3390/catal9121053
Submission received: 31 October 2019 / Revised: 3 December 2019 / Accepted: 9 December 2019 / Published: 11 December 2019
(This article belongs to the Special Issue Coordination Chemistry and Catalysis)

Abstract

:
The mononuclear Cu(II) complex [Cu((kNNO-HL)(H2O)2] (1) was synthesized using N-acetylpyrazine-2-carbohydrazide (H2L) and characterized by elemental analysis, IR spectroscopy, ESI-MS and single crystal X-ray crystallography. Two Fe(III) complexes derived from the same ligand viz, mononuclear [Fe((kNNO-HL)Cl2] (2) and the binuclear [Fe(kNNO-HL)Cl(μ-OMe)]2 (3) (synthesized as reported earlier), were also used in this study. The catalytic activity of these three complexes (13) was examined towards the oxidation of alcohols using tert-butyl hydroperoxide (TBHP) as oxidising agent under solvent-free microwave irradiation conditions. Primary and secondary benzyl alcohols (benzyl alcohol and 1-phenylethanol), and secondary aliphatic alcohols (cyclohexanol) were used as model substrates for this study. A comparison of their catalytic efficiency was performed. Complex 1 exhibited the highest activity in the presence of TEMPO as promoter for the oxidation of 1-phenylethanol with a maximum yield of 91.3% of acetophenone.

Graphical Abstract

1. Introduction

Transition metal complexes play a crucial role as catalysts or catalytic precursors in various organic transformations. The designing of new metal complexes with suitable multidentate ligands is a challenging research area aiming at finding better catalytic efficiency. Several transition metal complexes have already exhibited high catalytic activity in different organic oxidative transformations, such as oxidation of alkanes or alcohols, epoxidation, carboxylation, hydrocarboxylation, in C–C couplings, CO2 reduction, etc. [1,2,3,4,5,6,7,8,9,10,11,12,13,14]. The efficiency of metal complexes in catalysis is highly dependent on their structural configuration and electronic properties [10,11,15,16,17,18]. The ligands also play a supporting role towards the promising catalytic performance of the metal complexes. Therefore, selection of appropriate ligands for the synthesis of new metal complex catalysts is crucial.
In this study, we mainly focused on the selective oxidation of alcohols to the corresponding carbonyl compounds [18,19,20,21]. The products of oxidation of alcohols such as aldehydes and ketones are building blocks for many organic compounds [22,23,24,25] and show diverse applications, such as pharmaceuticals, agrochemicals, fragrances, fine chemicals and polymers [26,27]. The commencement of clean synthetic catalytic routes can be implemented by using green oxidants, green solvents or solvent free processes, energy saving beneficial techniques and involves the design of efficient catalysts from environmental and economic perspectives [28,29,30]. Based on the nature of the catalyst precursors, both homogeneous and heterogeneous (supported) catalysts have been reported by us towards oxidative catalytic reactions [3,4,5,6,7,8,9,29,30,31,32,33,34,35,36,37,38,39,40]. Although many catalysts have been developed to overcome ecological drawbacks, namely those using cheap and abundant metals, like copper and iron, metal catalytic systems operating under sustainable conditions are still challenging.
Considering all the above mentioned points herein we present the synthesis of a new mononuclear Cu(II) complex, [Cu((kNNO-HL)(H2O)2] (1), derived from the N-acetylpyrazine-2-carbohydrazide (H2L) ligand. Two Fe(III) complexes, the mononuclear [Fe((kNNO-HL)Cl2] (2) and the binuclear [Fe(kNNO-HL)Cl(μ-OMe)]2 (3), have also been synthesized using the same ligand as reported earlier [41]. The three complexes (13) were screened as catalytic precursors to assess their catalytic performances in the oxidation of alcohols. In the present investigation, we chose primary or secondary benzyl alcohols (benzyl alcohol, 1-phenyl alcohol) and secondary aliphatic alcohols (cyclohexanol) as model substrates and tert-butyl hydroperoxide (TBHP) as oxidising agent under microwave irradiation. MW irradiation acts as an alternative technique due to its simplicity and energy saving characteristics [6], usually leading to higher product yields and selectivities. In this investigation, the best catalytic conditions were optimized by comparing several cofactors like temperature, reaction time, influence of the presence of additives, etc.

2. Results and Discussion

2.1. Synthesis and Characterizations

The pro-ligand N-acetylpyrazine-2-carbohydrazide (H2L) was used to synthesize the mononuclear Cu(II) complex [Cu(kNNO-HL)(H2O)2] (1) (Scheme 1). The Fe(III) complexes [Fe(kNN′O-HL)Cl2] (2) and [Fe(kNN′O-HL)Cl(µ-OMe)]2 (3) derived from the same pro-ligand were obtained by a method from the literature [41]. As previously observed, in this case, the N-acetylpyrazine-2-carbohydrazide also acts as a mononegative tridentate N,N,O donor (HL) towards the metal centre [41] and shows different coordination behaviour from a related one derived from N-acetylsalicylhydrazide [42,43,44]. In this case, the –CONH– group attached to the pyrazine ring only deprotonates and favours the formation of two stable five-member chelate rings around the metal centre. In the case of N-acetylsalicylhydrazide, the ligand prefers to coordinate simultaneously with two metal ions in its two chelate pockets [42,43,44].
Characterizations of 13 were carried out by elemental analysis, spectroscopic (IR spectroscopy and ESI-MS) methods, and single-crystal X-ray diffraction techniques. In addition to the other characteristic stretching bands of the ligand, a blue shift was observed for the C=O stretching frequency at 1662 cm−1 (which is coordinated to the Cu(II) centre) and the presence of nitrate ion at 1384 cm−1 [33]. The m/z value of 1 indicates the loss of non-coordinate nitrate ion present in 1 (see Experimental).
The catalytic properties of 13 were investigated towards neat microwave assisted oxidation of alcohols (benzyl alcohol, 1-phenylethanol and cyclohexanol), and their catalytic activities were compared.

2.2. General Description of the Crystal Structure

Green single crystals suitable for the X-ray analysis of 1 (Supplementary Materials: CCDC 1962610) were obtained from methanol upon slow evaporation (in open air) at room temperature. The molecular structure of 1 obtained by X-ray analysis is presented in Figure 1. The crystallographic data and other processing parameters are summarized in Table 1, and selected dimensions (bond lengths and angles) are provided in Table 2.
Complex 1 crystallizes in the monoclinic C2/c space group and its asymmetric unit comprises the Cu(II) cation with one coordinated ligand, two water molecules, and a non-coordinate nitrate anion. The copper centre of 1 is five-coordinated with two nitrogen atoms (amido and pyrazine) and one keto oxygen atom from the HL ligand and two water molecules. The metal cation in 1 exhibits N2O3 coordination environment assuming a distorted square pyramid geometry (τ5 = 0.144), where τ5 = 0 for square pyramid and τ5 = 1 for trigonal bipyramid geometries [45]. The tridentate HL ligand and one water molecule occupy the square plane (N1N9O12O14). The axial position is occupied by another water molecule (O15). The central Cu(II) ion is located 0.214 Å above from the above-mentioned square plane towards apical oxygen O15.
Extensive inter- and intra-molecular hydrogen bond interactions were found in 1. A hydrogen bond 2D network (Figure 2) results from the contacts involving the water ligands, amine nitrogen of the ligand and nitrate anions.

2.3. Catalytic Studies

Complexes 13 were tested as catalyst precursors for the neat microwave (MW)-assisted oxidation of primary or secondary benzyl alcohols (benzyl alcohol, 1-phenylethanol) and secondary aliphatic alcohols (cyclohexanol) to the corresponding aldehydes (for primary alcohols), or ketones (for secondary alcohols), using aqueous tert-butyl hydroperoxide (TBHP) as oxidizing agent, under low-power (5–10 W) MW irradiation (Scheme 2). A high selectivity (towards the formation of acetophenone, benzaldehyde and cyclohexanone from their corresponding 1-phenylethanol, benzyl alcohol and cyclohexanol) was observed from the MW-assisted transformations, since no traces of by-products were detected by GC-MS analysis of the final reaction mixtures (only the unreacted alcohol was found, apart from the products). All three complexes of Cu(II) and Fe(III) (13) act as homogeneous catalyst precursors in neat conditions.
The influence of temperature and time was optimized using 1-phenylethanol as a model substrate. With a temperature increase from 80 °C to 120 °C, the catalytic performance of 13 was enhanced, exhibiting the highest activity (51.4% yield of acetophenone, entry 3, Table 2) after 30 min, with a corresponding TON (moles of product per mol of catalyst precursor) value of 512 in the presence of 1 under MW irradiation in the absence of any solvent (Figure 3). Analysing the yields of the peroxidative selective oxidation of 1-phenylethanol by complexes 13, it can clearly be seen that the catalytic activity of mononuclear Cu(II) complex 1 is higher than the other two mono and dinuclear Fe(III) complexes (2 and 3).
It was found that 1 h of MW irradiation provided the best catalytic conditions for yielding acetophenone, reaching a maximum in the presence of 13 and then slightly decreasing, conceivably due to overoxidation (Figure 4). Complexes 1, 2 and 3 exhibited maximum yields of 86.4%, 77.7% and 65.2% [Table 3, entries 4, 23 and 33, respectively].
The other heating technique, the conventional heating (oil bath) mode, was applied under the same optimized reaction conditions on catalytic system 1 to compare the product yield obtained by the MW-assisted selective oxidation of the chosen model substrate, 1-phenylethanol. From the conventional heating, using same conditions, 67.6% yield of acetophenone was achieved after 1 h, whereas 86.4% of acetophenone was obtained with MW irradiation. This clearly indicates the credibility of microwave irradiation as a beneficial technique. After 4 h of heating, the yield enhanced and reached from 67.6% to 87.2% [Table 3, entries 7 and 8]. Therefore, microwave irradiation reduces the reaction time required to achieve similar yields to one quarter of those obtained under conventional heating, which requires more energy.
The influence of various additives was explored in this study. Cu(II) complex 1 was chosen as a benchmark catalyst and its performance investigated, in the presence of nitric acid (HNO3), 2-pyrazinecarboxylic acid (HPCA), 2,2,6,6-tetramethylpiperidine-1-oxyl radical (TEMPO) and an oxygen-radical trap such as diphenyl amine (Ph2NH) towards neat microwave-assisted peroxidative oxidation of 1-phenylethanol at optimized conditions (Figure 5). In the presence of HNO3, the yield decreased dramatically from 86.4% to 25.6% [Table 3, entry 9]. The heteroaromatic acid HPCA also led to lower yields of acetophenone (51.8%) [Table 3, entry 8]. A highly favourable effect of additive was observed in presence of TEMPO for the MW-oxidation of 1-phenylethanol by 1–3 [Table 3, entries 10, 26 and 36]. The highest yield of 91.3% [TON (TOF) value of 457 (457)] was obtained for the Cu(II) complex 1, whereas Fe(III) complexes 2 and 3 accounted for 83.8% and 71.3%, respectively.
In contrast, a strong inhibitor effect of the catalytic activity of 13 was observed [Table 3, entries 11, 15 and 27] for the reactions carried out in the presence of Ph2NH. The addition of Ph2NH to the reaction mixture, a well-known oxygen radical trap [46], resulted in a significant yield drop compared to the reaction carried out under the same conditions but in the absence of such a radical trap. This suggests that oxygen radicals are generated during the reaction, which are trapped by the radical scavenger. The mechanism may involve the coordination of 1-phenylethanol followed by metal-centred dehydrogenation and oxidation of the alcohol through hydrogen abstraction or one-electron oxidation processes [47,48].
Since the best catalytic performances were obtained in the presence of 13 at 120 °C and 1 h MW irradiation, the oxidation of other alcohol substrates was tested under the same optimized reaction parameters. Oxidation of aromatic primary alcohol (benzyl alcohol) and aliphatic alcohol (cyclohexanol) using aq. tert-butyl hydroperoxide (ButOOH, TBHP, 70% aq. solution) as oxidant in the presence of catalyst precursors 13 under neat conditions yielded in lower value than secondary alcohol (1-phenylethanol) (Figure 6). In presence of 1, we found 33.4% of benzaldehyde after 1 h at 120 °C which increased up to maximum 41.2% yield in presence of TEMPO [Table 3, entries 12 and 14]. In the presence of 1, using optimized conditions, the oxidation of cyclohexanol, in the presence of TEMPO, exhibited the highest yield (67.9%) of cyclohexanone [Table 3, entry 18]. The oxidation of alcohol by TEMPO-free Cu(II) catalysts is also reported in the literature [49]. From the yield analysis of benzaldehyde and cyclohexanone for 13 catalytic systems, it was found that the Cu(II) complex acts as a more efficient catalyst precursor than Fe(III) complexes 2 and 3.
The isolated yield was determined by column chromatography using a mixture of ethylacetate and n-hexane (1:3) as eluent, and the purity of the product was verified by 1H NMR (SI Figures S1–S3). The catalytic reactions were performed with 10 mmol of each substrate (1-phenyl ethanol, benzyl alcohol and cyclohexanol) under optimized conditions and the isolated yield has been determined from the reaction products. The isolated yields found were 83.6% (for 1-phenyl ethanol to acetophenone), 32.3% (for benzyl alcohol to benzophenone) and 59.7% (for cyclohexanol to cyclohexanone), respectively. The isolated yields were found to be almost 8–9% lower than the yield obtained by GC-MS. This is probably due to the loss of product during the process of column chromatography.
For comparative purposes, the starting salts used in the synthesis of the complexes 13 were also tested [Table 3, entries 41–45].
The catalytic performance of 1 was also compared with some recent literature reports (Table 4) [50,51,52,53]. It is clear from Table 4 that complex 1 exhibits a good catalytic efficiency (a maximum yield of 91.3% was achieved in 1 h) in comparison to the reported ones.

3. Materials and Methods

Synthesis of the pro-ligand and metal complexes for this study was performed in open air. Reagents and solvents were used as commercially received, without further purification or drying. Cu(NO3)2·2.5H2O was used as metal precursor for the synthesis of complex 1. Elemental analyses (C, H and N) were carried out by the Microanalytical Service of the Instituto Superior Técnico. Bruker Vertex 70 instrument (Bruker Corporation, Ettlingen, Germany) was used for Infrared spectra (4000–400 cm−1) analysis in KBr pellets; wavenumbers are in cm−1. The 1H NMR spectrum of the ligand was recorded on a Bruker Avance II + 400.13 MHz (UltraShieldTM Magnet, Rheinstetten, Germany) spectrometer at room temperature. The internal reference was tetramethylsilane and the chemical shifts are reported in ppm in the 1H NMR spectrum. Mass spectra were recorded in a Varian 500-MS LC Ion Trap Mass Spectrometer (Agilent Technologies, Amstelveen, The Netherlands) equipped with an electrospray (ESI) ion source. The electrospray ionization was carried out with a flow rate and a drying gas optimized according to the particular sample with 35 p.s.i. nebulizer pressure. Scanning was performed in methanol solution from m/z 100 to 1200. The samples were analysed in the positive mode (capillary voltage = 80–105 V).

3.1. Synthesis of the Pro-Ligand H2L

The pro-ligand N-acetylpyrazine-2-carbohydrazide (H2L) was prepared according to the literature [41] upon acetylation of the pyrazine-2-carbohydrazide.
Yield: 84.0%. Anal. calc. for C7H8N4O2: C, 46.67; H, 4.48; N, 31.10. Found: C, 46.61; H, 4.53; N, 31.08. IR (KBr pellet, cm−1): 3336 ν(NH), 3223 ν(NH), 1698 ν(C=O), 1672 ν(C=O). 1H NMR (DMSO-d6, δ): 9.16–8.87 (m, 3H, C4H3N2), 8.75 (s, 2H, NH), 1.92 (s, 3H, CH3).

3.2. Synthesis of [Cu(kNN’O-HL)(H2O)2] (1)

The pro-ligand H2L (0.368 g, 1.00 mmol) was dissolved in 20 mL methanol solution and 0.245 g, 1.05 mmol Cu(NO3)2·2.5H2O was added to it. The reaction mixture was stirred for 30 min at a temperature of 50 °C. The resultant dark green solution was filtered, and the filtrate was kept in open air for crystallization. Green single crystals were isolated after 2 days, suitable for X-ray diffraction analysis.
Yield: 0.245 g (72%, with respect to Cu). Anal. Calcd for C7H11CuN5O7 (1): C, 24.67; H, 3.25; N, 20.55. IR (KBr pellet, cm−1): 3126 ν(NH), 1384 ν(ΝO3)-, 1700 ν(C=O), 1662 ν(C=O), 1034 ν(N–N). ESI-MS(+): m/z 278 [M-(NO3)]+ (100%).

3.3. Synthesis of [Fe(kNN’O-HL)Cl2] (2) and [Fe(kNN’O-HL)Cl(µ-OMe)]2 (3)

The mononuclear 2 and dinuclear Fe(III) complexes were synthesized as described in the literature [41].
[Fe(kNN’O-HL)Cl2] (2): Yield 70%, 0.21 g. Anal. Calcd for C7H7Cl2FeN4O2: C, 27.48; H, 2.31; N, 18.31. Found: C, 27.43; H, 2.28; N, 18.28. IR (KBr pellet, cm−1): 3130 ν(NH), 1702 ν(C=O), 1666 ν(C=O), 1037 ν(N–N). ESI-MS(+): m/z 306 [M+H]+ (100%).
[Fe(kNN’O-HL)Cl(µ-OMe)]2 (3): Yield 66 %, 0.19 g. Anal. Calcd for C16H20Cl2Fe2N8O6: C, 31.87; H, 3.34; N, 18.58. Found: C, 31.82; H, 3.32; N, 18.53. IR (KBr; cm−1): 3127 ν(NH), 1664 ν(C=O), 1638 ν(C=O), 1036 ν(N–N). ESI-MS(+): m/z 604 [M+H]+ (100%).

3.4. X-Ray Measurements

A single crystal of complex 1 of appropriate quality for X-ray diffraction analysis was chosen and immersed in cryo-oil, mounted in Nylon loops and measured at 297 K. Intensity data were collected using a Bruker AXS PHOTON 100 diffractometer with graphite monochromated Mo-Kα (λ 0.71073) radiation. Data collections were recorded using omega scans of 0.5° per frame and full sphere of data were obtained. Cell parameters were retrieved using Bruker SMART [54] software, and the data were refined using Bruker SAINT [54] on all the observed reflections. Absorption corrections were applied using SADABS [54]. Structures were solved by direct methods by using SIR97 [55] and refined with SHELXL2014 [56]. Calculations were performed using WinGX v2014.1 [57]. Those H-atoms bonded to carbon were included in the model at geometrically calculated positions and refined using a riding model. Uiso(H) were defined as 1.2Ueq of the parent carbon atoms for aromatic residues and 1.5Ueq for the methyl groups. The other hydrogen atoms (N–H) were in calculated positions as aromatic located in the difference Fourier synthesis and refined. Least square refinements with anisotropic thermal motion parameters were applied for all the non-hydrogen atoms and isotropic for the remaining atoms.

3.5. Catalytic Studies

The catalytic experiments were carried out under microwave irradiation in a focused microwave Anton Paar Monowave 300 (Anton Paar GmbH, Graz, Austria) discover reactor fitted with a rotational system and an IR temperature detector. 10 mL capacity cylindrical Pyrex tubes with a 13 mm internal diameter were used. Gas chromatographic (GC) measurements were carried out using a FISONS Instruments GC 8000 series gas chromatograph (Agilent Technologies, Santa Clara, CA, USA) with a DB-624 (J&W) capillary column (FID detector) and the Jasco-Borwin v.1.50 software (Jasco, Tokyo, Japan). The temperature of injection was 240 °C. The initial temperature, 120 °C, was maintained for 1 min, then raised 10 °C/min to 200 °C and held at 200 °C for 1 min. The carrier gas used was helium. GC-MS analyses were conducted using a Perkin-Elmer Clarus 600 C (Shelton, CT, USA) instrument (with He as the carrier gas) with an ionization voltage of 70 eV and a SGE BPX5 column (30 m × 0.25 mm × 0.25 µm). The comparison of the products retention times with those of known reference compounds enabled their identification. Moreover, their mass spectra to fragmentation patterns were compared with those obtained from the NIST spectral library of the computer software of the spectrometer.

Typical Procedures for the Catalytic Oxidation of Alcohols and Product Analysis

The oxidation reactions of the alcohol substrates were performed in the above-mentioned Pyrex tubes under focused microwave irradiation as follows: alcohol (5 mmol), catalyst precursor 13 (10 µmol, 0.2 mol% vs. substrate) and a 70% aqueous solution of tBuOOH (10 mmol) were introduced in the tube. This was then placed in the microwave reactor and the system was stirred and irradiated (5–10 W) for 0.5–2 h at 80–120 °C. After the reaction, the mixture was allowed to cool down to room temperature. In the case of 1-phenylethanol and cyclohexanol, 300 µL of benzaldehyde (internal standard) and 5 mL of NCMe (to extract the substrate and the organic products from the reaction mixture) were added. For benzyl alcohol, 90 µL cycloheptanone (internal standard) was added and 10 mL of diethyl ether (to extract the substrate and the organic products from the reaction mixture) was added. The mixture was stirred for 10 min. Then, a 1 µL sample was taken from the organic layer and analysed by gas chromatography. The product quantification used the internal standard method. The performed blank experiments indicated that only traces (<0.7%) of ketones (cyclohexanone or acetophenone or) are formed in a catalyst-free system. Flash column chromatography was performed on silica gel 60, 63–200 microns from Panreac; ethyl acetate/hexane as eluent to obtain the desired product.

4. Conclusions

In this work we successfully explored the catalytic activities of a mononuclear Cu(II) complex [Cu(kNNO-HL)(H2O)2] (1) and two Fe(III) complexes mononuclear [Fe(kNNO-HL)Cl2] (2) and the binuclear [Fe(kNNO-HL)Cl(μ-OMe)]2 (3) towards the oxidation of alcohols using tert-butyl hydroperoxide as oxidising agent under solvent-free microwave irradiation conditions. Three different alcohol substrates (benzyl alcohol, 1-phenylethanol and cyclohexanol) were used for this study to compare the catalytic performance of the catalytic precursors 13. All the catalyst precursors led to very good yields and selectivity towards the oxidation of alcohols. Complex 1 was found to be more efficient than the other two Fe(III) complexes, and the promotor TEMPO had an accelerating effect on the catalytic oxidation of alcohols. The regioselectivity of the catalytic system and the use of neat systems under low power MW irradiation were significant factors for energy saving and the sustainability of greener environments; therefore, the catalytic system could be useful for the future.

Supplementary Materials

CCDC 1962610 for 1 contains the supplementary crystallographic data for this paper. This data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif. Supplementary data to this article can be found online at https://www.mdpi.com/2073-4344/9/12/1053/s1.

Author Contributions

Conceptualization, M.S.; datacuration, T.R.B.; formalanalysis, T.R.B.; fundingacquisition, A.J.L.P. and L.M.D.R.S.M.; investigation, M.S.; methodology, M.S. and T.R.B.; project administration, A.J.L.P. and L.M.D.R.S.M.; resources, A.J.L.P.; software, M.S.; supervision, M.S. and L.M.D.R.S.M.; visualization, L.M.D.R.S.M.; writing—original draft, M.S. and T.R.B.; writing—review and editing, A.J.L.P. and L.M.D.R.S.M.

Funding

Authors are grateful to the Fundação para a Ciência e a Tecnologia: FCT (projects UID/QUI/00100/2019, PTDC/QEQ-ERQ/1648/2014 and PTDC/QEQ-QIN/3967/2014), Portugal, for financial support.

Acknowledgments

M.S. acknowledges the FCT and IST for a working contract “DL/57/2017” (Contract no. IST-ID/102/2018). The authors are thankful to the Portuguese NMR Network (IST-UL Centre) for access to the NMR facility and the IST Node of the Portuguese Network of mass-spectrometry for the ESI-MS measurements.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Gupta, K.C.; Sutar, A.K. Catalytic activities of Schiff base transition metal complexes. Coord. Chem. Rev. 2008, 252, 1420–1450. [Google Scholar] [CrossRef]
  2. Chelucci, G. Ruthenium and osmium complexes in C-C bond-forming reactions by borrowing hydrogen catalysis. Coord. Chem. Rev. 2017, 331, 1–36. [Google Scholar] [CrossRef]
  3. 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, 301–302, 200–239. [Google Scholar] [CrossRef]
  4. Sutradhar, M.; Roy Barman, T.; Pombeiro, A.J.L.; Martins, L.M.D.R.S. Ni(II)-Aroylhydrazone Complexes as Catalyst Precursors Towards Efficient Solvent-Free Nitroaldol Condensation Reaction. Catalysts 2019, 9, 554. [Google Scholar] [CrossRef] [Green Version]
  5. Sutradhar, M.; Kirillova, M.V.; Guedes da Silva, M.F.C.; Cai-Ming, L.; Pombeiro, A.J.L. Tautomeric effect of hydrazone Schiff bases in tetranuclear Cu(II) complexes: Magnetism and catalytic activity towards mild hydrocarboxylation of alkanes. Dalton. Trans. 2013, 42, 16578–16587. [Google Scholar] [CrossRef] [PubMed]
  6. Sutradhar, M.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Oxido vanadium complexes with tridentate aroylhydrazone as catalyst precursors for solvent-free microwave-assisted oxidation of alcohol. Appl. Cat. A Gen. 2015, 493, 50–57. [Google Scholar] [CrossRef]
  7. 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]
  8. Sutradhar, M.; Da Silva, M.F.C.G.; Pombeiro, A.J. A new cyclic binuclear Ni(II) complex as a catalyst towards nitroaldol (Henry) reaction. Catal. Commun. 2014, 57, 103–106. [Google Scholar] [CrossRef]
  9. Sutradhar, M.; Alegria, E.C.B.A.; Roy Barman, T.; 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]
  10. Martins, L.M.D.R.S.; Pombeiro, A.J.L. Tris(pyrazol-1yl)methane metal complexes for catalytic mild oxidative functionalizations of alkanes, alkenes and ketones. Coord. Chem. Rev. 2014, 265, 74–88. [Google Scholar] [CrossRef]
  11. Martins, L.M.D.R.S. C-scorpionate complexes: Ever young catalytic tools. Coord. Chem. Rev. 2019, 396, 89–102. [Google Scholar] [CrossRef]
  12. Ribeiro, A.P.C.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Carbon dioxide-to-methanol single-pot conversion using a C-scorpionate iron(II) catalyst. Green Chem. 2017, 19, 4801–4962. [Google Scholar] [CrossRef]
  13. 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]
  14. Montoya, C.A.; Gómez, C.F.; Paninho, A.B.; Nunes, A.V.M.; Mahmudov, K.T.; Najdanovic-Visak, V.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Nunes da Ponte, M.; Pombeiro, A.J.L. Cyclic carbonate synthesis from CO2 and epoxides using zinc(II) complexes of arylhydrazones of β-diketones. J. Catal. 2016, 335, 135–140. [Google Scholar] [CrossRef] [Green Version]
  15. Abrantes, M.; Valente, A.; Pillinger, M.; Goncalves, I.S.; Rocha, J.; Romao, C.C. Organotin–Oxometalate Coordination Polymers as Catalysts for the Epoxidation of Olefins. J. Catal. 2002, 209, 237–244. [Google Scholar] [CrossRef]
  16. Togni, A.; Venanzi, L.M. Nitrogen Donors in Organometallic Chemistry and Homogeneous Catalysis. Angew. Chem. Int. Ed. Engl. 1994, 33, 497–526. [Google Scholar] [CrossRef]
  17. Fache, F.; Dunjic, B.; Gamez, P.; Lemaire, M. Recent advances in homogeneous and heterogeneous asymmetric catalysis with nitrogen-containing ligands. Top. Catal. 1997, 4, 201–209. [Google Scholar] [CrossRef]
  18. Marko, I.E.; Giles, P.R.; Tsukazaki, M.; Brown, S.M. Copper-Catalyzed Oxidation of Alcohols to Aldehydes and Ketones: An Efficient, Aerobic Alternative. Science 1996, 274, 2044–2046. [Google Scholar] [CrossRef] [Green Version]
  19. Larock, R.C. Comprehensive Organic Transformations; Wiley-VCH: New York, NY, USA, 1999. [Google Scholar]
  20. Wang, X.; Wu, G.; Wei, W.; Sun, Y. Solvent-free oxidation of alcohols by hydrogen peroxide over chromium Schiff base complexes immobilized on MCM-41. Trans. Met. Chem. 2010, 35, 213–220. [Google Scholar] [CrossRef]
  21. Farhadi, S.; Zaidi, M. Polyoxometalate–zirconia (POM/ZrO2) nanocomposite prepared by sol-gel process: A green and recyclable photocatalyst for efficient and selective aerobic oxidation of alcohols into aldehydes and ketones. Appl. Catal. A 2009, 354, 119–126. [Google Scholar] [CrossRef]
  22. Brühne, F.; Wright, E. Benzaldehyde, Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2011; Volume 5, pp. 223–233. [Google Scholar]
  23. Lee, D.G.; Spitzer, U.A. Aqueous dichromate oxidation of primary alcohols. J. Org. Chem. 1970, 35, 3589–3590. [Google Scholar] [CrossRef]
  24. Prabhakaran, P.V.; Venkatachalam, S.; Ninan, K.N. Permanganate ion supported over crosslinked polyvinylamine as an oxidising agent for alcohols. Eur. Poly. J. 1999, 35, 1743–1746. [Google Scholar] [CrossRef]
  25. Abad, A.; Almela, C.; Corma, A.; García, H. Efficient chemoselective alcohol oxidation using oxygen as oxidant. Superior performance of gold over palladium catalysts. Tetrahedron 2006, 62, 6666–6672. [Google Scholar] [CrossRef]
  26. Tojo, G.; Fernandez, M. Oxidation of Alcohols to Aldehydes and Ketones: A Guide to Current Common Practice; Springer: New York, NY, USA, 2006. [Google Scholar]
  27. Ullmann’s Encyclopedia of Industrial Chemistry, 6th ed.; Wiley-VCH: Weinheim, Germany, 2002.
  28. Mardani, H.R.; Golchoubian, H. Effective oxidation of benzylic and aliphatic alcohols with hydrogen peroxide catalyzed by a manganese(III) Schiff-base complex under solvent-free conditions. Tetrahedron Lett. 2006, 47, 2349–2352. [Google Scholar] [CrossRef]
  29. Sutradhar, M.; Martins, L.M.; Roy Barman, 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]
  30. 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]
  31. 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]
  32. 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]
  33. 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]
  34. Sutradhar, M.; Alegria, E.C.B.A.; 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]
  35. 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] [PubMed] [Green Version]
  36. 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] [PubMed] [Green Version]
  37. 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]
  38. Alexandru, M.; Cazacu, M.; Arvinte, A.; Shova, S.; Turta, C.; Simionescu, B.C.; Dobrov, A.; Alegria, E.C.B.A.; Martins, L.M.D.R.S.; Pombeiro, A.J.L.; et al. Chlorido-bridged dimanganese(II) complexes of the Schiff base derived from [2+2]condensation of 2,6-diformyl-4-methylphenol and 1,3-bis(3-aminopropyl)tetramethyldisiloxane: Structure, Magnetism, Electrochemical Behaviour and Catalytic Oxidation of Secondary Alcohols. Eur. J. Inorg. Chem. 2014, 2014, 120–131. [Google Scholar]
  39. 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. C. R. Chim. 2015, 18, 758–765. [Google Scholar] [CrossRef]
  40. Silva, T.F.S.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.; Kuznetsov, M.L.; Fernandes, A.R.; Silva, A.; Santos, S.; Pan, C.-J.; Lee, J.-F.; Hwang, B.-J.; et al. Cobalt complexes with pyrazole ligands as catalysts for the peroxidative oxidation of cyclohexane. XAS studies and biological applications. Chem. Asian J. 2014, 9, 1132–1143. [Google Scholar] [CrossRef]
  41. 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-carbohydrazide-Fe(III)-chloro Complexes. ChemistrySelect 2018, 3, 8349–8355. [Google Scholar] [CrossRef]
  42. Sutradhar, M.; Kirillova, M.V.; Guedes da Silva, M.F.C.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. A hexanuclear oxovanadium(IV,V) complex bearing N,O-ligand as a highly efficient alkane oxidation catalyst. Inorg. Chem. 2012, 51, 11229–11231. [Google Scholar] [CrossRef] [Green Version]
  43. 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]
  44. 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]
  45. Addison, W.; Rao, T.N.; Reedijk, J.; van Rijn, J.; Verschoor, G.C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc. Dalton Trans. 1984, 1349–1356. [Google Scholar] [CrossRef]
  46. Moiseeva, I.N.; Gekham, A.E.; Minin, V.V.; Larin, G.M.; Bashtanov, M.E.; Krasnovskii, A.A.; Moiseev, I.I. Free radical/singlet dioxygen system under the conditions of catalytic hydrogen peroxide decomposition. Kinet. Catal. 2000, 41, 170–177. [Google Scholar] [CrossRef]
  47. Figiel, P.J.; Leskelä, M.; Repo, T. TEMPO-Copper(II) Diimine-Catalysed Oxidation of Benzylic Alcohols in Aqueous Media. Adv. Synth. Catal. 2007, 349, 1173–1179. [Google Scholar] [CrossRef]
  48. Ahmad, J.U.; Figiel, P.J.; Räisänen, M.T.; Leskelä, M.; Repo, T. Aerobic oxidation of benzylic alcohols with bis(3,5-di-tert-butylsalicylaldimine)copper(II) complexes. Appl. Catal. A 2009, 371, 17–21. [Google Scholar] [CrossRef]
  49. Xu, B.; Lumb, J.-P.; Arndtsen, B.A. A TEMPO-Free Copper-Catalyzed Aerobic Oxidation of Alcohols. Angew. Chem. Int. Ed. 2015, 54, 1–5. [Google Scholar]
  50. Gao, J.; Ren, Z.-G.; Lang, J.-P. Oxidation of benzyl alcohols to benzaldehydes in water catalyzed by a Cu(II) complex with a zwitterionic calix[4]arene ligand. J. Organometall. Chem. 2015, 792, 88–92. [Google Scholar] [CrossRef]
  51. Jlassia, R.; Ribeiro, A.P.C.; Mendes, M.; Rekik, W.; Tiago, G.A.O.; Mahmudov, K.T.; Naïli, H.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Arylhydrazone Cd(II) and Cu(II) complexes as catalysts for secondary alcohol oxidation. Polyhedron 2017, 129, 182–188. [Google Scholar] [CrossRef]
  52. Zhao, Y.; Yu, C.; Wu, S.; Zhang, W.; Xue, W.; Zeng, Z. Synthesis of Benzaldehyde and Benzoic Acid by Selective Oxidation of Benzyl Alcohol with Iron(III) Tosylate and Hydrogen Peroxide: A Solvent-Controlled Reaction. Catal. Lett. 2018, 148, 3082–3092. [Google Scholar] [CrossRef]
  53. Pramanick, R.; Bhattacharjee, R.; Sengupta, D.; Datta, A.; Goswami, S. An Azoaromatic Ligand as Four Electron Four Proton Reservoir: Catalytic Dehydrogenation of Alcohols by Its Zinc(II) Complex. Inorg. Chem. 2018, 57, 6816–6824. [Google Scholar] [CrossRef]
  54. Bruker, A.X.S. APEX2 & SAINT.; AXS Inc.: Madison, WI, USA, 2004. [Google Scholar]
  55. Altomare, A.; Burla, M.C.; Camalli, M.; Cascarano, G.L.; Giacovazzo, C.; Guagliardi, A.; Moliterni, A.G.G.; Polidori, G.; Spagna, R. SIR97: A New Tool for Crystal Structure Determination and Refinement. J. Appl. Cryst. 1999, 32, 115–119. [Google Scholar] [CrossRef]
  56. Sheldrick, G.M. A Short History of SHELX. Acta Crystallogr. Sect. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Farrugia, L.J. WinGX and ORTEP for Windows: An Update. J. Appl. Cryst. 2012, 45, 849–854. [Google Scholar] [CrossRef]
Scheme 1. Syntheses of [Cu((kNNO-HL)(H2O)2] (1), [Fe(kNN′O-HL)Cl2] (2) and [Fe(kNN′O-HL)Cl(µ-OMe)]2 (3).
Scheme 1. Syntheses of [Cu((kNNO-HL)(H2O)2] (1), [Fe(kNN′O-HL)Cl2] (2) and [Fe(kNN′O-HL)Cl(µ-OMe)]2 (3).
Catalysts 09 01053 sch001
Figure 1. Structural representation (obtained by single crystal X-ray analysis) of 1 with partial atom labelling scheme. H-bond interactions are represented in dotted lines in light blue colour. Symmetry codes for generating equivalent atoms: (i) 1.5 − x,1/2 + y,1/2 − z; (ii) 1 − x,1 − y,1 − z; (iii) 2 − x, y,1/2 − z; (iv) x,1 − y,1/2 + z.
Figure 1. Structural representation (obtained by single crystal X-ray analysis) of 1 with partial atom labelling scheme. H-bond interactions are represented in dotted lines in light blue colour. Symmetry codes for generating equivalent atoms: (i) 1.5 − x,1/2 + y,1/2 − z; (ii) 1 − x,1 − y,1 − z; (iii) 2 − x, y,1/2 − z; (iv) x,1 − y,1/2 + z.
Catalysts 09 01053 g001
Figure 2. Hydrogen-bonded 2D network of 1, viewed along the crystallographic b axis. The nitrate ions are drawn by the space-filled model.
Figure 2. Hydrogen-bonded 2D network of 1, viewed along the crystallographic b axis. The nitrate ions are drawn by the space-filled model.
Catalysts 09 01053 g002
Scheme 2. Microwave-assisted neat oxidation of 1-phenylethanol, benzyl alcohol and cyclohexanol to acetophenone, benzaldehyde and cyclohexanone, respectively, in presence of catalyst precursors (13) using aq. tert-butyl hydroperoxide (ButOOH, TBHP, 70% aq. solution) as oxidant.
Scheme 2. Microwave-assisted neat oxidation of 1-phenylethanol, benzyl alcohol and cyclohexanol to acetophenone, benzaldehyde and cyclohexanone, respectively, in presence of catalyst precursors (13) using aq. tert-butyl hydroperoxide (ButOOH, TBHP, 70% aq. solution) as oxidant.
Catalysts 09 01053 sch002
Figure 3. Dependence on the temperature for MW-assisted neat oxidation of 1-phenylethanol using 13 as catalyst precursors. Reaction conditions: 5 mmol of substrate, 10 µmol (0.2 mol% vs. substrate) of 13, 10 mmol of TBHP (2 eq., 70% in H2O), 30 min reaction time, MW irradiation (5–10 W).
Figure 3. Dependence on the temperature for MW-assisted neat oxidation of 1-phenylethanol using 13 as catalyst precursors. Reaction conditions: 5 mmol of substrate, 10 µmol (0.2 mol% vs. substrate) of 13, 10 mmol of TBHP (2 eq., 70% in H2O), 30 min reaction time, MW irradiation (5–10 W).
Catalysts 09 01053 g003
Figure 4. Yield analysis of MW-assisted neat peroxidative oxidation of 1-phenylethanol using 13 as catalyst precursors with respect to time. Reaction conditions: 5 mmol of substrate, 10 µmol (0.2 mol% vs. substrate) of 13, 10 mmol of TBHP (2 eq., 70% in H2O), 120 °C, MW irradiation (5–10 W).
Figure 4. Yield analysis of MW-assisted neat peroxidative oxidation of 1-phenylethanol using 13 as catalyst precursors with respect to time. Reaction conditions: 5 mmol of substrate, 10 µmol (0.2 mol% vs. substrate) of 13, 10 mmol of TBHP (2 eq., 70% in H2O), 120 °C, MW irradiation (5–10 W).
Catalysts 09 01053 g004
Figure 5. Influence of different additives on the yield of acetophenone, obtained from MW-assisted neat peroxidative oxidation of 1-phenylethanol in presence of catalyst precursor 1. Reaction conditions: 5 mmol of substrate, 10 µmol (0.2 mol% vs. substrate) of 1, 10 mmol of TBHP (2 eq., 70% in H2O), additives [n(additive)/n(catalyst) = 25], 120 °C, MW irradiation (5–10 W).
Figure 5. Influence of different additives on the yield of acetophenone, obtained from MW-assisted neat peroxidative oxidation of 1-phenylethanol in presence of catalyst precursor 1. Reaction conditions: 5 mmol of substrate, 10 µmol (0.2 mol% vs. substrate) of 1, 10 mmol of TBHP (2 eq., 70% in H2O), additives [n(additive)/n(catalyst) = 25], 120 °C, MW irradiation (5–10 W).
Catalysts 09 01053 g005
Figure 6. Yield analysis of MW-assisted neat peroxidative oxidation of 1-phenylethanol, benzyl alcohol and cyclohexanol in presence of catalyst precursors 13. Reaction conditions: 5 mmol of substrate, 10 µmol (0.2 mol% vs. substrate) of 13, 10 mmol of TBHP (2 eq., 70% in H2O), 120 °C, 1 h reaction time, MW irradiation (5–10 W).
Figure 6. Yield analysis of MW-assisted neat peroxidative oxidation of 1-phenylethanol, benzyl alcohol and cyclohexanol in presence of catalyst precursors 13. Reaction conditions: 5 mmol of substrate, 10 µmol (0.2 mol% vs. substrate) of 13, 10 mmol of TBHP (2 eq., 70% in H2O), 120 °C, 1 h reaction time, MW irradiation (5–10 W).
Catalysts 09 01053 g006
Table 1. Crystal data and structure refinement details for complex 1.
Table 1. Crystal data and structure refinement details for complex 1.
1
Empirical formulaC7H11CuN5O7
Formula Weight340.75
Crystal systemMonoclinic
Space groupC2/c
Temperature/K297 (2)
a13.8799 (9)
b16.9433 (10)
c11.9900 (7)
α/°90
β/°117.763 (2)
γ/°90
V3)2495.1 (3)
Z8
Dcalc (g cm−3)1.814
μ(Mo Kα) (mm−1)1.79
Rfls. collected/unique/observed15921/2292/1997
Rint0.029
Final R1a, wR2b (I ≥ 2σ)0.027, 0.072
Goodness-of-fit on F21.07
aR = Σ||Fo| − |Fc||/Σ|Fo|; b wR(F2) = [Σw(|Fo|2 − |Fc|2)2/Σw|Fo|4]½
Table 2. Selected bond distances (Å) and angles (°) in complex 1.
Table 2. Selected bond distances (Å) and angles (°) in complex 1.
Cu1—N91.8900 (18)
Cu1—O141.9066 (17)
Cu1—O122.0111 (15)
Cu1—N12.0813 (18)
Cu1—O152.260 (2)
N9—Cu1—O14166.20 (10)
N9—Cu1—O1280.33 (7)
O14—Cu1—O1294.01 (7)
N9—Cu1—N179.43 (7)
O14—Cu1—N1103.48 (8)
O12—Cu1—N1157.54 (7)
N9—Cu1—O1597.24 (8)
O14—Cu1—O1596.11 (9)
O12—Cu1—O1599.49 (7)
N1—Cu1—O1592.63 (7)
Table 3. Data a for the selective peroxidative oxidation of cyclohexane with TBHP (70% aq.) using complexes 13 as catalyst precursors.
Table 3. Data a for the selective peroxidative oxidation of cyclohexane with TBHP (70% aq.) using complexes 13 as catalyst precursors.
EntryCatalystSubstrateTemperature (°C)Reaction Time (h)AdditiveYield (%) bTON (TOF (h-1)) c
111-phenyl ethanol800.5-32.3162 (324)
21000.5-44.8224 (448)
31200.5-51.4257 (514)
41201.0-86.4432 (432)
51201.5-69.8349 (233)
61202.0-65.4327 (164)
7 d1201-67.6338 (338)
8 d1204 87.2436 (109)
8 e1201HPCA51.8259 (259)
9 f1201HNO325.6128 (128)
10 g1201TEMPO91.3457 (457)
11 h1201Ph2NH5.930 (30)
12benzyl alcohol1201-33.4167 (167)
13 d1201-19.598 (98)
14 g1201TEMPO41.2206 (206)
15 h1201Ph2NH3.618 (18)
16cyclohexanol1201-65.6328 (328)
17 d1201-23.9120 (120)
18 g1201TEMPO67.9274 (274)
19 h1201Ph2NH3.920 (20)
2021-phenyl ethanol800.5-20.1101 (101)
211000.5-24.6123 (123)
221200.5-46.2193 (385)
231201.0-77.7389 (389)
241201.5-69.3347 (231)
251202.0-66.9335 (167)
26 g1201.0TEMPO83.8419 (419)
27 h1201.0Ph2NH4.724 (24)
28benzyl alcohol1201.0-26.7133 (133)
29cyclohexanol1201.0-35.8 179 (179)
3031-phenyl ethanol800.5-11.256 (112)
311000.5-20.3102 (204)
321200.5-28.6143 (286)
331201.0-65.2326 (326)
341201.5-60.4302 (201)
351202.0-56.8284 (142)
36 g1201.0TEMPO71.3377 (377)
37 h1201.0Ph2NH2.814 (14)
38benzyl alcohol1201.0-18.995 (95)
39cyclohexanol1201.0-36.7184 (184)
40Cu(NO3).2.5 H2O1-phenyl ethanol1201.0-6.231 (31)
41benzyl alcohol1201.0-2.714 (14)
42cyclohexanol1201.0-4.322 (22)
43FeCl31-phenyl ethanol1201.0-4.925 (25)
44benzyl alcohol1201.0-2.211 (11)
45cyclohexanol1201.0-3.417 (17)
a Reaction conditions: 5 mmol of substrate, 10 µmol (0.2 mol% vs. substrate) of catalyst precursor 13, 10 mmol of TBHP (2 eq., 70% in H2O), MW irradiation (5 W). b Moles of ketone product per 100 moles of alcohol. c Turnover number = number of moles of product per mol of catalyst precursor; TOF = TON per hour (values in brackets). d Conventional heating. e n(HPCA)/n(catalyst) = 25. f n(HNO3)/n(catalyst) = 25. g n(TEMPO)/n(catalyst) = 25. h n(Ph2NH)/n(catalyst) = 25. MW irradiation (5–10 W).
Table 4. Comparison of catalytic activity of 1 with other known compounds.
Table 4. Comparison of catalytic activity of 1 with other known compounds.
CatalystAmount (mol%)SubstrateOxidant Temp (°C)Time (h)Yield (%)Ref
[Cu(II)L1(H2O)]I20.25Benzyl alcoholH2O2 (in the presence of TEMPO and K2CO3)60249951
[Cu(im)(µ-HL2-1κO:2κNOO′)]20.151-phenylethanolH2O2 (in the presence of TEMPO)80 (MW irradiation)17452
Fe(OTs)3·6H2O0.1Benzyl alcoholH2O2601685.653
ZnL3Cl25Benzyl alcoholO2 (in the presence of KtBuO, Zn dust)6024 7654
[Cu(kNNO-HL)(H2O)2]0.2Benzyl alcoholTBHP (in the presence of TEMPO)120 (MW irradiation)191.3This work
H4L1 = 5,11,17,23-tetrakis(trimethylammonium)-25,26,27,28-tetrahydroxycalix[4]arene. HL2 = 2-[2-(2,4-dioxopentan-3-ylidene)hydrazinyl]terephthalic acid. TsOH = Tosyl alcohol. L3 = 2,6-bis(phenylazo)pyridine.

Share and Cite

MDPI and ACS Style

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 Neat Microwave Assisted Oxidation of Alcohols. Catalysts 2019, 9, 1053. https://doi.org/10.3390/catal9121053

AMA Style

Sutradhar M, Roy Barman T, Pombeiro AJL, Martins LMDRS. Cu(II) and Fe(III) Complexes Derived from N-Acetylpyrazine-2-Carbohydrazide as Efficient Catalysts Towards Neat Microwave Assisted Oxidation of Alcohols. Catalysts. 2019; 9(12):1053. https://doi.org/10.3390/catal9121053

Chicago/Turabian Style

Sutradhar, Manas, Tannistha Roy Barman, Armando J. L. Pombeiro, and Luísa M. D. R. S. Martins. 2019. "Cu(II) and Fe(III) Complexes Derived from N-Acetylpyrazine-2-Carbohydrazide as Efficient Catalysts Towards Neat Microwave Assisted Oxidation of Alcohols" Catalysts 9, no. 12: 1053. https://doi.org/10.3390/catal9121053

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