Next Article in Journal
γ-Alumina Nanoparticle Catalyzed Efficient Synthesis of Highly Substituted Imidazoles
Next Article in Special Issue
A Copper-Based Metal-Organic Framework as an Efficient and Reusable Heterogeneous Catalyst for Ullmann and Goldberg Type C–N Coupling Reactions
Previous Article in Journal
Enhanced Visible Light Photocatalytic Activity of Br-Doped Bismuth Oxide Formate Nanosheets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Greener Selective Cycloalkane Oxidations with Hydrogen Peroxide Catalyzed by Copper-5-(4-pyridyl)tetrazolate Metal-Organic Frameworks

1
Chemical Engineering Departament, Instituto Superior de Engenharia de Lisboa, Instituto Politécnico de Lisboa, R. Conselheiro Emídio Navarro, 1959-007 Lisboa, Portugal
2
Centro de Química Estrutural, Complexo I, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisboa, Portugal
3
Department of Chemistry, School of Basic Sciences, Indian Institute of Technology Indore, IET-DAVV Campus, Khandwa Road, Indore 452017, India
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2015, 20(10), 19203-19220; https://doi.org/10.3390/molecules201019203
Submission received: 25 September 2015 / Revised: 13 October 2015 / Accepted: 16 October 2015 / Published: 21 October 2015
(This article belongs to the Special Issue Metal Mediated Activation of Small Molecules)

Abstract

:
Microwave assisted synthesis of the Cu(I) compound [Cu(µ4-4-ptz)]n [1, 4-ptz = 5-(4-pyridyl)tetrazolate] has been performed by employing a relatively easy method and within a shorter period of time compared to its sister compounds. The syntheses of the Cu(II) compounds [Cu33-4-ptz)42-N3)2(DMF)2]n∙(DMF)2n (2) and [Cu(µ2-4-ptz)2(H2O)2]n (3) using a similar method were reported previously by us. MOFs 1-3 revealed high catalytic activity toward oxidation of cyclic alkanes (cyclopentane, -hexane and -octane) with aqueous hydrogen peroxide, under very mild conditions (at room temperature), without any added solvent or additive. The most efficient system (2/H2O2) showed, for the oxidation of cyclohexane, a turnover number (TON) of 396 (TOF of 40 h−1), with an overall product yield (cyclohexanol and cyclohexanone) of 40% relative to the substrate. Moreover, the heterogeneous catalytic systems 13 allowed an easy catalyst recovery and reuse, at least for four consecutive cycles, maintaining ca. 90% of the initial high activity and concomitant high selectivity.

Graphical Abstract

1. Introduction

In the last decade, 5-substituted tetrazole ligands have been evolving as one of the more useful linkers for generation of functional materials due to their interesting chemical and structural properties. A few methods have been followed until now for in-situ tetrazole generation by cycloaddition between an organonitrile and an azide in the presence of a transition metal ion, where the solvothermal process has dominated to a large extent over the other ones [1,2,3]. Although there are a few reports where tetrazoles had been prepared under mild conditions in the presence of metal ions and other catalysts [4,5,6,7], descriptions of construction of a MOF containing 5-(4-pyridyl)tetrazolate (4-ptz) building blocks in a controlled manner other than those involving a solvothermal process (at lower temperatures and pressures via a similar cycloaddition path) are scarce. In the present study, a highly improved synthesis of 1 is reported (Scheme 1), whereas the syntheses of 2 and 3 (Scheme 2) had previously been reported by us [8]. These compounds were generated while studying the effect of reaction conditions on the coordination modes of 4-pytz by employing the [2 + 3] cycloaddition as a tool for generating the 5-substituted tetrazole ligands in-situ from 4-pyridinecarbonitrile and NaN3 in the presence of a copper(II) salt. Curiously, though there is a report with a structure similar to 1 (which was produced via the solvothermal method in the time period of one day) [3], in our case the time has been significantly reduced to 10 min by using microwave irradiation. This procedure can be considered as an easy to handle and relatively energy efficient method for the production of Cu(I) MOFs in pure form. It is interesting to note that, to our knowledge, this is the first procedure where a MOF is generated by a one-pot reduction of Cu(II) to Cu(I) and formation of a tetrazolate linker via [2 + 3] cycloaddition within such a short time (10 min).
Scheme 1. Synthesis of MOF compound [Cu(µ4-4-ptz)]n (1).
Scheme 1. Synthesis of MOF compound [Cu(µ4-4-ptz)]n (1).
Molecules 20 19203 g006
Although copper based metal-organic materials as catalysts for hydrocarbon oxidation have been studied in the last few years with diverse ligand systems [9,10,11,12,13,14,15,16,17,18], the application of tetrazole based copper-organic frameworks for the oxidation of alkanes under mild oxidation conditions remains largely unexplored.
Scheme 2. Synthetic procedures of compounds [Cu33-4-ptz)42-N3)2(DMF)2]n∙(DMF)2n (2) and {[Cu(µ2-4-ptz)2(H2O)2]}n (3) [8].
Scheme 2. Synthetic procedures of compounds [Cu33-4-ptz)42-N3)2(DMF)2]n∙(DMF)2n (2) and {[Cu(µ2-4-ptz)2(H2O)2]}n (3) [8].
Molecules 20 19203 g007
Selective partial oxidation of alkanes is an important topic with potential in terms of economic and ecological perspectives of sustainable chemistry. However, efficient catalytic oxidation of alkanes still remains as a challenging topic. A relevant example of selective alkane oxidation [9,19,20,21,22,23,24,25,26,27] with industrial significance concerns the oxidation of cyclohexane to cyclohexanol and cyclohexanone that are important reagents for the production of adipic acid and caprolactam used for the manufacture of nylon [19,20,21,27]. The current industrial route uses a homogeneous cobalt species as catalyst, dioxygen as oxidant and requires considerably harsh conditions (150 °C), forming the oxidation products in low yields (ca. 5%) to achieve a good selectivity (ca. 85%), [19,20,21]. The development of more efficient catalysts, under milder conditions, at room temperature and using low toxicity media and oxidizing agents is needed [18,19,20,21,27,28,29,30,31,32,33,34]. Hydrogen peroxide is one of the best options in this regard since H2O is the sole by-product and exhibits an atom efficiency and e-factor similar to dioxygen [34]. Thus, peroxidative (with H2O2) alkane oxidations were the selected reactions for the present study. Herein we report the catalytic performances of Cu(I) 1 and Cu(II) 2 and 3 compounds toward the oxidation of cyclic alkanes (cyclopentane, -hexane and -octane) under very mild and green (solvent- and additive-free) conditions as a significant step towards the protection of environment and quality of life.

2. Results and Discussion

2.1. Synthesis and Spectroscopic Characterization of [Cu(µ4-4-ptz)]n (1)

Compound 1 was prepared by microwave irradiation, at 130 °C for 10 min, of a mixture containing copper(II) chloride, sodium azide and 4-cyanopyridine in 1:2:4 molar ratios using a water-DMF mixture (1:6, v:v), from which the yellow colored Cu(I) compound 1 crystallized upon cooling to room temperature.
The obtained compound was characterized by IR spectroscopy, elemental analysis and single-crystal X-ray crystallography. The IR spectrum of 1 shows a strong band at 1650 cm−1 [35,36] indicating the presence of the tetrazolate moiety. The SCXRD data of compound 1 (Figure 1) revealed that its structural properties are similar to those of the reported compound [3]. Powder X-ray diffraction (PXRD) patterns recorded for bulk samples of 1 (Figure 1) show a very good matching with the respective simulated patterns (acquired from the single crystal X-ray data) thus demonstrating their phase purity.
Figure 1. PXRD patterns of 1, simulated (green) and bulk sample (red).
Figure 1. PXRD patterns of 1, simulated (green) and bulk sample (red).
Molecules 20 19203 g001
Single crystal X-ray diffraction studies indicate that 1 had grown as a rigid three-dimensional framework with one-dimensional open-ended pores. In this framework Cu(I) is acting as the central metal ion. It is coordinated by four nitrogen atoms of four 4-ptz ligands, among them one atom being the pyridine nitrogen (N1) and the remaining three atoms pertaining to the tetrazolate ring (N2, N3 and N5) from four 4-ptz linkers, and exhibits a distorted tetrahedron geometry (Figure 2a). Each 4-ptz is coordinated to four different metal centers and is acting as a tetradentate ligand.
Figure 2. Structural fragments of 1 representing (a) the basic unit; (b) a view of rhomboid voids along the a-axis. Color codes: Cu brown, C black, and N blue.
Figure 2. Structural fragments of 1 representing (a) the basic unit; (b) a view of rhomboid voids along the a-axis. Color codes: Cu brown, C black, and N blue.
Molecules 20 19203 g002

2.2. Catalytic Oxidation of Cycloalkanes

Complexes 13 were tested as catalysts for the oxidation of cyclic alkanes (cyclopentane, -hexane and -octane) to the corresponding alcohol and ketone (the final products) mixtures via formation of cycloalkyl hydroperoxide (CyOOH, primary product) [31,37,38,39,40], according to Scheme 3 and Table 1 and Table 2 (see below). The catalytic systems are based on any of the above Cu(I) 1 or Cu(II) 2 or 3 complexes, hydrogen peroxide (30% aqueous solution) as the oxidizing agent, at room temperature (r.t.), in the absence of added solvent or additives. The use of other environmentally benign [34] peroxidative oxidants, such as tert-butyl hydroperoxide (TBHP, 70% aqueous solution), was also considered.
Scheme 3. Solvent-free oxidation of cycloalkanes (n = 1, 2 or 4) to the corresponding alcohol and ketone mixtures.
Scheme 3. Solvent-free oxidation of cycloalkanes (n = 1, 2 or 4) to the corresponding alcohol and ketone mixtures.
Molecules 20 19203 g008
Table 1. Oxidation of selected cycloalkanes using 13 as catalysts (selected data) a.
Table 1. Oxidation of selected cycloalkanes using 13 as catalysts (selected data) a.
EntryCatalystSubstrateOxidantTON [TOF (h−1)] bOLYield (%) c ONETotal dOL/ONE Ratio
11cyclopentaneH2O2260 (26)14.311.726.01.2
2TBHP440 (44)5.516.722.20.3
3cyclohexaneH2O2369 (37)18.818.136.91.0
4TBHP664 (66)11.321.933.20.5
5cyclooctaneH2O2218 (22)11.710.121.81.2
6TBHP434 (43)8.313.421.70.6
72cyclopentaneH2O2276 (28)13.114.527.60.9
8TBHP416 (42)7.613.220.80.6
9cyclohexaneH2O2396 (40)20.818.839.61.1
10TBHP704 (70)12.922.335.20.6
11cyclooctaneH2O2343 (34)21.412.934.31.7
12TBHP620 (62)5.325.731.00.2
133cyclopentaneH2O2215 (22)11.99.621.51.2
14TBHP218 (22)4.06.910.90.6
15cyclohexaneH2O2366 (37)16.520.136.60.8
16TBHP710 (71)13.921.635.50.6
17cyclooctaneH2O2285 (29)17.810.728.51.7
18TBHP468 (47)8.714.723.40.6
19nonecyclopentaneH2O2-2.11.13.21.9
20TBHP-1.11.82.90.6
21cyclohexaneH2O2-2.41.43.81.7
22TBHP-1.32.53.80.5
23cyclooctaneH2O2-2.21.53.71.5
24TBHP-0.71.3 2.00.5
a Reaction conditions unless stated otherwise: 5.0 mmol of substrate, 2.5–5 μmol of catalyst, 10.0 mmol of oxidant, r.t., 10 h reaction time. Yield and TON determined by GC analysis (upon treatment with PPh3). b Turnover number = number of moles of products per mol of catalyst; TOF = TON per hour (values in brackets). c Molar yield (%) based on substrate, i.e., moles of product (alcohol (OL) or ketone (ONE)) per 100 mol of cycloalkane. d Moles of alcohol + ketone per 100 moles of cyclohexane.
The formation of CyOOH (under the conditions of Table 1) is proved by using the method proposed by Shul’pin [37,38,39,40]. The addition of PPh3 prior to the GC analysis of the products results in a marked increase of the amount of alcohol (due to reduction of CyOOH by PPh3, with formation of phosphane oxide) and a corresponding decrease of ketone, as observed in other catalytic systems [9,10,11,12,13,41,42,43,44,45].
Table 2. Peroxidative oxidation of cyclohexane with H2O2 (selected data) a.
Table 2. Peroxidative oxidation of cyclohexane with H2O2 (selected data) a.
EntryCatalystn(cat.)/n(CyH) × 103n(H2O2)/n(cat.) × 10−3Reaction Time (h)Yield (%) bTON [TOF (h−1)] c
OLONETotal d
1140.51011.16.717.845 (4.5)
2211016.611.528.1141 (14)
31.31.51016.217.834.0262 (26)
40.82.51012.421.934.3429 (43)
50.45102.118.019.9498 (50)
6120.255.52.68.181 (8.1)
7120.58.42.611.0110 (11)
81219.15.814.9149 (15)
9122.510.511.321.8218 (22)
1012517.113.730.8308 (31)
111212.518.716.034.7347 (35)
1212246.019.325.3253 (25)
13 e12101.31.93.232 (3.2)
14 f12102.82.35.151 (5.1)
15 g12101.10.21.313 (1.3)
16 h12101.61.12.727 (2.7)
17 i12109.325.034.3343 (34)
18240.51010.25.415.639 (4)
19211017.610.828.4142 (14)
201.31.51016.720.337.0285 (29)
210.82.51016.919.536.4455 (46)
220.45105.8 10.716.5413 (41)
23120.257.52.49.999 (9.9)
24120.510.63.313.9139 (14)
2512114.44.919.3193 (19)
26122.519.78.828.5285 (29)
2712517.818.135.9359 (36)
281212.518.019.637.6376 (38)
2912245.521.727.2272 (27)
30 e12100.60.71.313 (1.3)
31 f12101.13.54.545 (4.5)
32 g12101.70.82.525 (2.5)
33 h12100.91.22.121 (2.1)
34 i121011.926.037.9379 (38)
35340.51010.28.718.947 (4.7)
36211017.112.930.0150 (15)
371.31.51016.318.835.1270 (27)
380.82.51015.220.135.3441 (44)
390.45108.415.323.7593 (59)
40120.257.14.411.5115 (12)
41120.58.35.714.0140 (14)
421219.18.517.6176 (18)
43122.515.611.927.5275 (28)
4412519.514.133.6336 (34)
451212.522.920.235.1351 (35)
46212414.940.123.5235 (24)
47 e12100.80.61.414 (1.4)
48 f12101.21.12.323 (2.3)
49 g12101.40.72.121 (2.1)
50 h12100.71.21.919 (1.9)
51 i12108.426.935.3353 (35)
a Reaction conditions (unless stated otherwise): cyclohexane (5.0 mmol), 2–20 μmol of 13, H2O2 (10 mmol), r.t., 0.25–24 h reaction time. Percentage of yield, TON determined by GC analysis (upon treatment with PPh3). b Molar yield (%) based on substrate, i.e., moles of products (cyclohexanol (OL) or cyclohexanone (ONE)) per 100 mol of cyclohexane. c Turnover number = moles of products per mol of catalyst; TOF = TON per hour (values in brackets). d Moles of cyclohexanol + cyclohexanone per 100 moles of cyclohexane. e Reaction in the presence of nitric acid. f Reaction in the presence of Hpca. g Reaction in the presence of CBrCl3 (5.0 mmol). h Reaction in the presence of Ph2NH (5.0 mmol). i values from GC analysis prior to addition of PPh3 (for comparative purposes).
Complex 2 provides the most efficient catalytic system, achieving overall yields (relative to the alkane) up to 39.6%, 34.3% and 27.6% for the peroxidative (with H2O2) oxidation of cyclohexane, -octane and -pentane, respectively (Table 1, entries 9, 11 and 7, respectively) after 10 h reaction time (Figure 3) which can conceivably be due to the presence of azide ligands. Their basic character could promote proton-transfer steps, a feature that is favorable to the occurrence of oxidation catalysis with peroxides. The highest yields were obtained for cyclohexane oxidation (39.6%, 36.9% and 33.6% for 2, 1 and 3, entries 9, 3 and 15, Table 1, respectively).
In general, the need of lower catalyst loads (0.05 mol% vs. substrate) and, thus, higher turnover numbers (TONs, number of moles of products per mol of catalyst) were found using TBHP (up to 710, Table 1). Moreover, TBHP appears to favor the formation of ketone (ratios alcohol/ketone equal or lower than 0.6, Table 1). The higher activity of this oxidant is also apparent on the over-oxidation products detected by GC-MS (mainly 1,2-cyclohexanediol and 1,4-cyclohexanedione) when the reaction was run in its presence.
Blank tests were performed under the same reaction conditions (Table 1) but in the absence of MOFs and no significant conversion of the cycloalkanes was observed (entries 19–24, Table 1). Moreover, the replacement of 13 by their precursor salt CuCl2·2H2O, resulted in a drastic decrease of activity (maximum total yields of 4, 5 and 5%, respectively for cyclopentane, -hexane and -octane), suggesting a significant role of the ligands in the catalytic oxidation of the tested alkanes.
Figure 3. Dependence of the overall yield (mol %, based on substrate) of the products (cyclohexanol + cyclohexanone) on the reaction time, for the oxidation of cyclohexane. Reaction conditions: cyclohexane (5.0 mmol), 5.0 μmol of 1 (•), 2 (♦) or 3 (■), n(H2O2)/n(catalyst) (2 × 103), r.t.
Figure 3. Dependence of the overall yield (mol %, based on substrate) of the products (cyclohexanol + cyclohexanone) on the reaction time, for the oxidation of cyclohexane. Reaction conditions: cyclohexane (5.0 mmol), 5.0 μmol of 1 (•), 2 (♦) or 3 (■), n(H2O2)/n(catalyst) (2 × 103), r.t.
Molecules 20 19203 g003
It should be emphasized that yields approaching 40% obtained herein for cyclohexane oxidation can be considered as remarkably high for the oxidation of very inert alkanes, much above those reported for other copper(II) complexes (e.g., bearing N,O-ligands such as aminopolyalcohols, scorpionates or derivatives [45,46,47], functionalized azo derivatives of β-diketones [17] or Schiff bases), although the values of the present work are obtained at higher reaction times (10 h instead of the usual 6 h). This can conceivably be due to a lower activity toward further oxidation of the alcohol/ketone mixture of 13 that avoid the over oxidation usually reported for the other Cu systems, or may result from a longer lifetime of our catalysts.
The obtained yield is also much higher than that of the industrial process [19,21,27] in spite of the used mild conditions [ambient temperature, atmospheric pressure, with an aqueous green oxidant, with considerable low loads of catalyst (up to 0.2 mol % of Cu catalyst vs. substrate) and without the addition of any solvent or additive].
Moreover, a high selectivity towards the formation of the alcohol/ketone mixtures is exhibited by our systems, since no traces of by-products were detected by GC-MS analysis of the final reaction mixtures for the optimized conditions. These features are of utmost importance for the establishment of a greener catalytic process for cyclohexane oxidation.
The influence of various reaction parameters, such as time, the amounts of catalyst and oxidant and presence of additives were investigated for the most active substrate (cyclohexane)/oxidant (H2O2) system and the results are summarized in Table 2 and Figure 3 and Figure 4.
The yield drop observed (Figure 3) for reaction times higher than 10 h results from the occurrence of subsequent reactions in the oxidative medium. Over-oxidation products such as 1,3-cyclohexanediol and 1,4-cyclohexanediol were detected by CG-MS for 24 h reaction time.
Figure 4. Dependence of the overall yield (mol %, based on substrate) of the products (cyclohexanol + cyclohexanone) on the amount of oxidant (H2O2, molar ratio relatively to 1 (•), 2 (♦) or 3 (■)) in the oxidation of cyclohexane. Reaction conditions: n(H2O2)/n(catalyst) (0–5 × 103), cyclohexane (5.0 mmol), r.t., 10 h.
Figure 4. Dependence of the overall yield (mol %, based on substrate) of the products (cyclohexanol + cyclohexanone) on the amount of oxidant (H2O2, molar ratio relatively to 1 (•), 2 (♦) or 3 (■)) in the oxidation of cyclohexane. Reaction conditions: n(H2O2)/n(catalyst) (0–5 × 103), cyclohexane (5.0 mmol), r.t., 10 h.
Molecules 20 19203 g004
The effect of the peroxide-to-catalyst molar ratio is depicted in Figure 4. The increase of the peroxide amount up to then (H2O2)/n(catalyst) molar ratio of 2 × 103 leads to the maximum products yields. Further increase of the oxidant amount results in a marked yield drop due to over oxidation reactions at higher H2O2 amounts. In fact, the adipic acid precursor 1,2-cyclohexanediol was detected by GC-MS using the conditions of entry 22, Table 2.
The previously recognized usual promoting effect of an inorganic acid [11,12,13,45,46,47,48,49,50,51,52,53,54] or pyrazinecarboxylic acid (Hpca) [43,51,52,53,54,55,56,57] on the peroxidative oxidation of alkanes catalyzed by homogeneous [27,31,32,43,48,51,52,54,56] or supported [51,52,53,54,56] metallic species is not observed for the present systems. Moreover, the presence of acid (either mineral or organic) has a strong inhibitor effect of the catalytic activity (Table 2, entries 13, 14, 30, 31, 47 or 47). A similar behavior was found for C-scorpionate Au(III) complexes [56].
As observed for other Cu and different metal catalytic systems [9,10,11,13,17,37,45,46,47,49,50,51,54,56,57], introduction of a radical trap (CBrCl3 or Ph2NH) into the reaction mixture results in a considerable suppression of the catalytic activity. This behavior, along with the formation of cyclohexyl hydroperoxide (typical primary product in radical-type cyclohexane oxidation) supports a free-radical mechanism in this study.
As previously proposed for several homogeneous or heterogeneous Mn+1/n (e.g., V, Re, Fe, Cu or Au) catalytic systems [43,48,52,54,56,57] we can propose the following mechanism: copper-catalyzed decomposition of the peroxide ROOH (R = H or tBu) leads to the oxygen-centered radicals ROO and RO, upon oxidation by Cu(II) or reduction by Cu(I), respectively (reactions 1 and 2; in the case of Cu(I) compound 1, reactions 1 and 2 occur in the reverse order, i.e., first 2 and then 1). Water is believed to catalyze H+-shift steps towards the formation of RO [58,59,60]. Cycloalkyl radical Cy is then formed upon H-abstraction from cycloalkane CyH by RO (reaction 3). Reaction of Cy with dioxygen leads to CyOO (reaction 4), and CyOOH can then be formed upon H-abstraction from ROOH by CyOO (reaction 5) or upon reduction of the latter to CyOO by Cu(I) followed by protonation. Metal-assisted decomposition of CyOOH to CyO and CyOO (reactions 6 and 7) would then lead to cyclohexanol (CyOH) and cyclohexanone (Cy-H=O) products (reactions 8 and 9) [60].
Cu(II) + ROOH → ROO + H+ + Cu(I)
Cu(I) + ROOH → RO + Cu(II) + HO
RO + CyH → ROH + Cy
Cy+ O2 → CyOO
CyOO + ROOH → CyOOH + ROO
CyOOH + Cu(I) → CyO + Cu(II) + HO
CyOOH + Cu(II) → CyOO + H+ + Cu(I)
CyO + CyH → CyOH + Cy
2CyOO → CyOH + Cy-H=O + O2
The initial availability of easily oxidized copper(I) to copper(II) species, or vice versa (easily reduced copper(II) to copper (I) species), to decompose the peroxide is crucial for this peroxidative oxidation, since the formation of the oxygen-centered radicals ROO and RO radicals is the key step for the occurrence of the C–H abstraction from the alkane.
Catalyst recyclability was investigated for up to four consecutive cycles for all the catalysts 13 on the oxidative media used for the conversion of cyclohexane. On completion of each stage, the products were analyzed as usually and the catalyst was recovered by filtration from the reaction mixture, thoroughly washed with acetonitrile and dried overnight at 60 °C. The subsequent cycle was initiated upon addition of new standard portions of all other reagents. The filtrate was analyzed relative to the presence of copper by atomic absorption spectroscopy and the hypothesis of catalyst leaching was excluded. Moreover, the filtrate was tested in a new reaction (by addition of fresh reagents), and no oxidation products were detected.
Figure 5 shows the recyclability of the systems: all were able to be reused while maintain almost the original level of activity after several consecutive reaction cycles (e.g., in a second, third and fourth run, the observed activity of 2 was 97%, 95%and 92% of the initial one) with a rather high selectivity to cyclohexanol and cyclohexanone.
Figure 5. Effect of the catalyst recycling on the overall yield of the products from the cyclohexane oxidation catalyzed by 13.
Figure 5. Effect of the catalyst recycling on the overall yield of the products from the cyclohexane oxidation catalyzed by 13.
Molecules 20 19203 g005
In addition, to demonstrate the structure conservation, catalyst 2 was analyzed by FT-IR before and after the catalytic reaction, and no significant changes were detected. This suggests that 2 is a true heterogeneous catalyst, and no catalytically active species were released into the solution.

3. Experimental Section

3.1. General Information

All chemical reagents used in the experiments were purchased from commercial sources and no further purification was employed before using them for reactions. Microwave assisted synthesis of complex 1 was achieved in a focused microwave Discover reactor (150 W, CEM, Buckingham, UK), using a reaction tube of 10 mL capacity with a 13 mm internal diameter, fitted with a rotational system and an IR temperature detector (CEM). Infrared spectra (4000–500 cm−1) were recorded with a Tensor 27 (with MIR source, Zn–Se beam splitter and DLaTGS detector, Bruker, Bremen, Germany) of samples in KBr pellets. Elemental analyses were carried out with a Thermo-Flash 2000 elemental analyzer (Thermo Scientific, Waltham, MA, USA). The spectrophotometric measurements were performed on a Cary 100 UV-Vis spectrophotometer (Varian, Santa Clara, CA, USA) using a quartz cuvette with a path length of 1 cm. Powder X-ray diffraction patterns for complex 1 were recorded on a Smart Lab X-ray diffractometer (Rigaku, Wilmingtons, MA, USA). The X-rays used were of wavelength of 0.154 nm (CuK-α) produced using a sealed tube and detected using a linear counting detector (Scintillator NaI photomultiplier detector).
X-ray crystallography: Single crystal X-ray structural studies of compound 1 were performed on a CCD (Oxford Diffraction, Agilent Technologies, Santa Clara, CA, USA) SUPER NOVA diffractometer. Data were collected at 150(2) K using graphite-monochromoated Mo Kα radiation (λα = 0.71073 Å). The strategy for the data collection was evaluated by using the CrysAlisProCCD software (Oxford Diffraction, Agilent Technologies). The data were collected by the standard phi-omega scan techniques and were scaled and reduced using CrysAlisPro RED software (Oxford Diffraction, Agilent Technologies). The structure was solved by direct methods using SHELXS-97 and refined by full matrix least-squares with SHELXL-97, refining on F2 [61,62]. The positions of all the atoms were obtained by direct methods. All non-hydrogen atoms were refined anisotropically. The remaining hydrogen atoms were placed in geometrically constrained positions and refined with isotropic temperature factors, generally 1.2Ueq of their parent atoms. The crystal and refinement data are summarized in Table 3.
Gas chromatographic (GC) measurements were carried out using a FISONS Instruments GC 8000 series gas chromatograph with a FID detector and a capillary column (DB-WAX, column length: 30 m; internal diameter: 0.32 mm, FISONS, Markham, ON, Canada) and the Jasco-Borwin software (version 1.50, FISONS). The temperature of injection was 240 °C. The initial temperature was maintained at 100 °C for 1 min, then raised 10 °C/min to 180 °C and held at this temperature for 1 min. Helium was used as the carrier gas. GC-MS analyses were performed using a Perkin Elmer Clarus 600 C instrument (He as the carrier gas, Linde, Lisboa, Portugal). The ionization voltage was 70 eV. Gas chromatography was conducted in the temperature-programming mode, using a SGE BPX5 column (30 m × 0.25 mm × 0.25 µm, FISONS). Reaction products were identified by comparison of their retention times with known reference compounds, and by comparing their mass spectra to fragmentation patterns obtained from the NIST spectral library stored in the computer software of the mass spectrometer.
Table 3. Crystallographic data and refinement details for 1.
Table 3. Crystallographic data and refinement details for 1.
1
Empirical formulaC6H4CuN5
Mr (g·mol−1)209.68
Crystal systemMonoclinic
Space groupP21/c
a (Å)5.8169(2)
b (Å)16.8804(6)
c (Å)9.0264(5)
α (°)90
β (°)94.070
γ (°)90
V3)884.08(7)
Z4
Dcalcd (mgm−3)1.575
F(000)416
GOF1.259
Reflections collected/unique5386/1546
Final R indices R1 = 0.0355, wR2 = 0.1032
R indices (all data)R1 = 0.0367, wR2 = 0.1036

3.2. Synthesis and Characterization of Complex [Cu(µ4-4-ptz)]n (1)

A greenish brown mixture of CuCl2∙2H2O (51 mg, 0.3 mmol), NaN3 (39 mg, 0.6 mmol) and 4-cyanopyridine (125 mg, 1.2 mmol) in H2O and DMF (1 mL:6 mL) mixture was placed in a reaction tube that was irradiated with microwave radiation for 10 min at 130 °C. Cooling the reaction mixture to room temperature resulted in deposition on the walls of reaction tube of yellow colored, single crystals X-ray analysis quality crystals, and a bulk micro crystalline sample that precipitated out, whose powder X-ray diffraction (PXRD) patterns exactly matched simulated patterns from single crystal X-ray data (Figure 2) confirming the purity of the bulk sample. Yield = 53%, anal calc. for C6H4CuN5: C, 34.37, H, 1.92, N, 33.40, found: C, 34.5, H, 1.98, N, 33.43. IR (KBr): 1650(s), 1625(s) 1558(w), 1434(m), 1390(m), 1210(m), 1109(m).

3.3. Typical Procedures for the Catalytic Oxidation of Cycloalkanes and Product Analysis

The peroxidative oxidation reactions were typically carried out as follows: 0.1–20 μmol of the catalyst was added to 5.00 mmol of the cycloalkane, whereafter 10.00 mmol of 30% H2O2 (1.02 mL) or of 70% TBHP (688 μL) were added and the reaction solution was stirred for 10 h at r.t. and normal pressure. In the experiments with radical traps, CBrCl3 (5.00 mmol) or NHPh2 (5.00 mmol) was added to the reaction mixture.
Catalyst recyclability was investigated, for up to four consecutive cycles. Each cycle was initiated after the preceding one upon addition of new typical portions of all other reagents. After completion of each run, the products were analyzed and the catalyst was recovered by filtration, washed with several portions of acetonitrile and dried in oven overnight at 60 °C.
The products analysis was performed as follows: 90 μL of cycloheptanone (internal standard), 10.00 mL of diethyl ether (to extract the substrate and the organic products from the reaction mixture) were added. The obtained mixture was stirred during 10 min and then a sample (1 μL) was taken from the organic phase and analyzed by gas chromatography (GC) by the internal standard method. Subsequently, an excess of solid triphenylphosphine was added to the final organic phase (to reduce the cyclohexyl hydroperoxide, if formed, to the corresponding alcohol, and hydrogen peroxide to water) and the mixture was analyzed again to estimate the amount of cyclohexyl hydroperoxide, following a method developed by Shul’pin [31,37,38,39]). For determination of oxygenate concentrations only data obtained after treatment of the reaction sample with PPh3 were used. Authentic samples of all oxygenated products were used to attribute the peaks in chromatograms. Blank tests indicate that no oxidation takes place in the absence of the Cu complex or the oxidant.

4. Conclusions

A novel microwave assisted methodology has been adopted to generate, within a few minutes, the Cu(I) based metal organic framework [Cu(µ4-4-ptz)]n (1). Copper based MOFs 13 act as catalysts for the mild and selective oxidation of cyclic alkanes in added solvent- and additive-free systems. A comparative study of their catalytic efficiency towards different cycloalkane substrates and oxidants has been performed. Furthermore, these heterogeneous greener catalytic systems allowed their easy recovery and reuse, at least for four consecutive cycles, maintaining over 90% of the initial activity and concomitant rather high selectivity. Moreover, the use of an aqueous medium at room temperature, without the requirement of an organic solvent, is a significant step towards the development of green catalytic systems in the field.

Supplementary Materials

CCDC940213 contains the supplementary crystallographic data for compound 1. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html, or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: (+44) 1223-336-033; or e-mail: [email protected].

Acknowledgments

We are grateful for the financial support received from the Council of Scientific and Industrial Research, New Delhi and from the Fundação para a Ciência e a Tecnologia (FCT), Portugal, and its project UID/QUI/00100/2013. One of us (M.S.) thanks CSIR for the award of JRF in a CSIR sponsored project. We are also thankful to Sophisticated Instrument Center, IIT Indore for the structure elucidation.

Author Contributions

RN, MS and SM (Shaikh Mobin) synthesized and characterized the MOF compounds 13; LM performed the catalytic studies; LM, AP and SM (Suman Mukhopadhyay) analyzed the data and wrote the paper. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, X.-L.; Li, N.; Tian, A.-X.; Ying, J.; Zhao, D.; Liu, X.-J. A hexa-nuclear {Cu6(ptz)6} cluster containing five [Cu2N4] units with cycle connecting cycle mode induced by tetrazole-based ligand to modify Keggin anions. Inorg. Chem. Commun. 2012, 25, 60–64. [Google Scholar] [CrossRef]
  2. Sha, J.-Q.; Sun, J.-W.; Li, M.-T.; Wang, C.; Li, G.-M.; Yan, P.-F.; Sun, L.-J. Secondary spacer modulated assembly of polyoxometalate based metal–organic frameworks. Dalton Trans. 2013, 42, 1667–1677. [Google Scholar] [CrossRef] [PubMed]
  3. Darling, K.; Ouellette, W.; Zubieta, J. Solid state coordination chemistry of copper with pyridyltetrazoles: Structural consequences of incorporation of coordinating anions. Inorg. Chim. Acta 2012, 392, 52–60. [Google Scholar] [CrossRef]
  4. Myznikov, L.V.; Roh, J.; Artamonova, T.V.; Hrabalek, A.; Koldobskii, G.I. Tetrazoles: LI. Synthesis of 5-Substituted Tetrazoles under Microwave Activation. J. Org. Chem. 2007, 43, 765–767. [Google Scholar] [CrossRef]
  5. Himo, F.; Demko, Z.P.; Noodleman, L.; Sharpless, K.B. Why Is Tetrazole Formation by Addition of Azide to Organic Nitriles Catalyzed by Zinc(II) Salts? J. Am. Chem. Soc. 2003, 125, 9983–9987. [Google Scholar] [CrossRef] [PubMed]
  6. Zhou, Y.; Yao, C.; Ni, R.; Yang, G. Amine Salt–Catalyzed Synthesis of 5-Substituted 1H-Tetrazoles from Nitriles. Synth. Commun. 2010, 40, 2624–2632. [Google Scholar] [CrossRef]
  7. Aureggi, V.; Sedelme, G. 1,3-Dipolar Cycloaddition: Click Chemistry for the Synthesis of 5-Substituted Tetrazoles from Organoaluminum Azides and Nitriles. Angew. Chem. Int. Ed. 2007, 46, 8440–8444. [Google Scholar] [CrossRef] [PubMed]
  8. Nasani, R.; Saha, M.; Mobin, S.M.; Martins, L.M.D.R.S.; Pombeiro, A.J.L.; Kirillov, A.M.; Mukhopadhyay, S. Copper-organic frameworks assembled from in-situ generated 5-(4-pyridyl)tetrazole building blocks: Synthesis, structural features, topological analysis and catalytic oxidation of alcohols. Dalton Trans. 2014, 43, 9944–9954. [Google Scholar] [CrossRef] [PubMed]
  9. Pombeiro, A.J.L. Advances in Organometallic Chemistry and Catalysis, The Silver/Gold Jubilee ICOMC Celebratory Book; John Wiley & Sons: Hoboken, NJ, USA, 2014. [Google Scholar]
  10. Kirillov, A.M.; Kirilova, M.V.; Pombeiro, A.J.L. Multicopper complexes and coordination polymers for mild oxidative functionalization of alkanes. Coord. Chem. Rev. 2012, 256, 2741–2759. [Google Scholar] [CrossRef]
  11. 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]
  12. 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]
  13. Di Nicola, C.; Garau, F.; Karabach, Y.Y.; Martins, L.M.D.R.S.; Monari, M.; Pandolfo, L.; Pettinari, C.; Pombeiro, A.J.L. Trinuclear Triangular Copper(II) Clusters. Synthesis, Electrochemical Studies and Catalytic Peroxidative Oxidation of Cycloalkanes. Eur. J. Inorg. Chem. 2009, 666–676. [Google Scholar] [CrossRef]
  14. 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]
  15. Conde, A.; Vilella, L.; Balcells, D.; Diaz-Requejo, M.; Lledos, A.; Perez, P. Introducing Copper as Catalyst for Oxidative Alkane Dehydrogenation. J. Am. Chem. Soc. 2013, 135, 3887–3896. [Google Scholar] [CrossRef] [PubMed]
  16. Perraud, O.; Sorokin, A.B.; Dutasta, J.-P.; Martinez, A. Oxidation of cycloalkanes by H2O2 using a copper–hemicryptophane complex as a catalyst. Chem. Commun. 2013, 49, 1288–1290. [Google Scholar] [CrossRef] [PubMed]
  17. Kopylovich, M.N.; Nunes, A.C.C.; Mahmudov, K.T.; Haukka, M.; Mac Leod, T.C.O.; Martins, L.M.D.R.S.; Kuznetsov, M.L.; Pombeiro, A.J.L. Complexes of copper(II) with 3-(ortho-substituted phenylhydrazo)pentane-2,4-diones: Syntheses, properties and catalytic activity for cyclohexane oxidation. Dalton Trans. 2011, 40, 2822–2836. [Google Scholar] [CrossRef] [PubMed]
  18. Hwang, Y.K.; Férey, G.; Lee, U.-H.; Chang, J.-S. Liquid Phase Oxidation of Organic Compounds by Metal-Organic Frameworks. In Liquid Phase Oxidation via Heterogeneous Catalysis, Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 2013; Chapter 8. [Google Scholar]
  19. Weissermel, K.; Arpe, H.J. Industrial Organic Chemistry, 2nd ed.; VCH Press: Weinheim, Germany, 1993. [Google Scholar]
  20. Clark, J.H.; Macquarrie, D.J. Handbook of Green Chemistry and Technology; Wiley: Hoboken, NJ, USA, 2002. [Google Scholar]
  21. Ullmann’s Encyclopedia of Industrial Chemistry, 6th ed.; Wiley-VCH: Weinheim, Germany, 2002.
  22. Crabtree, R.H. Organometallic alkane CH activation. J. Organomet. Chem. 2004, 689, 4083–4091. [Google Scholar] [CrossRef]
  23. 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]
  24. Shul’pin, G.B. Transition Metals for Organic Synthesis, 2nd ed.; Beller, M., Bolm, C., Eds.; Wiley-VCR: New York, NY, USA, 2004; Volume 2, p. 215. [Google Scholar]
  25. Derouane, E.G.; Parmon, V.; Lemos, F.; Râmoa Ribeiro, F. Sustainable Strategies for the Upgrading of Natural Gas: Fundamentals, Challenges, and Opportunities; Springer: Dordrecht, The Netherlands, 2005; Volume 191. [Google Scholar]
  26. Sheldon, R.A.; Arends, I.; Hanefeld, U. Green Chemistry and Catalysis; Wiley-VCH: Weinheim, Germany, 2007. [Google Scholar]
  27. Schuchardt, U.; Cardoso, D.; Sercheli, R.; Pereira, R.; Da Cruz, R.S.; Guerreiro, M.C.; Mandelli, D.; Spinacé, E.V.; Pires, E.L. Cyclohexane oxidation continues to be a challenge. Appl. Catal. A Gen. 2001, 211, 1–17. [Google Scholar] [CrossRef]
  28. 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]
  29. Shul’pin, G.B.; Attanasio, D.; Suber, L. Efficient H2O2 Oxidation of Alkanes and Arenes to Alkyl Peroxides and Phenols Catalyzed by the System Vanadate-Pyrazine-2-Carboxylic Acid. J. Catal. 1993, 142, 147–152. [Google Scholar] [CrossRef]
  30. Schuchardt, U.; Mandelli, D.; Shul’pin, G.B. Methyltrioxorhenium catalyzed oxidation of saturated and aromatic hydrocarbons by H2O2 in air. Tetrahedron Lett. 1996, 37, 6487–6490. [Google Scholar] [CrossRef]
  31. Shul’pin, G.B. Metal-catalyzed hydrocarbon oxygenations in solutions: The dramatic role of additives: A review. J. Mol. Catal. A Chem. 2002, 189, 39–66. [Google Scholar] [CrossRef]
  32. Shul’pin, G.B.; Nizova, G.V.; Kozlov, Y.N.; Cuervo, L.G.; Süss-Fink, G. Hydrogen Peroxide Oxygenation of Alkanes Including Methane and Ethane Catalyzed by Iron Complexes in Acetonitrile. Adv. Synth. Catal. 2004, 346, 317–332. [Google Scholar] [CrossRef]
  33. Anisia, K.S.; Kumar, A. Oxidation of cyclohexane with molecular oxygen using heterogeneous silica gel catalyst bonded with [1,2-bis(salicylidene amino)-phenylene] zirconium complex. Appl. Catal. A Gen. 2004, 273, 193–200. [Google Scholar] [CrossRef]
  34. Strukul, G.; Scarso, A. Environmentally Benign Oxidants. In Liquid Phase Oxidation via Heterogeneous Catalysis, Organic Synthesis and Industrial Applications; Clerici, M.G., Kholdeeva, O., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 2013; Chapter 1. [Google Scholar]
  35. Mukhopadhyay, S.; Lasri, J.; Charmier, M.A.J.; Silva, M.F.C.G.; Pombeiro, A.J.L. Microwave synthesis of mono- and bis-tetrazolato complexes via 1,3-dipolar cycloaddition of organonitriles with platinum(II)-bound azides. Dalton Trans. 2007, 5297–5304. [Google Scholar] [CrossRef]
  36. Smolenski, P.; Mukhopadhyay, S.; Guedes Silva, M.F.C.; Charmier, M.A.J.; Pombeiro, A.J.L. New water-soluble azido- and derived tetrazolato-platinum(II) complexes with PTA. Easy metal-mediated synthesis and isolation of 5-substituted tetrazoles. Dalton Trans. 2008, 6546–6555. [Google Scholar] [CrossRef] [PubMed]
  37. Shul’pin, G.B. Metal-catalysed hydrocarbon oxidations. C. R. Chim. 2003, 6, 163–178. [Google Scholar] [CrossRef]
  38. Shul’pin, G.B.; Kozlov, Y.N.; Shul’pina, L.S.; Kudinov, A.R.; Mandelli, D. Extremely Efficient Alkane Oxidation by a New Catalytic Reagent H2O2/Os3(CO)12/Pyridine. Inorg. Chem. 2009, 48, 10480–10482. [Google Scholar]
  39. Shul’pin, G.B.; Kozlov, Y.N.; Shul’pina, L.S.; Petrovskiy, P.V. Oxidation of alkanes and alcohols with hydrogen peroxide catalyzed by complex Os3(CO)10(µ-H)2. Appl. Organomet. Chem. 2010, 24, 464–472. [Google Scholar] [CrossRef]
  40. Shul’pin, G.B. C–H functionalization: thoroughly tuning ligands at a metal ion, a chemist can greatly enhance catalyst’s activity and selectivity. Dalton Trans. 2013, 42, 12794–12818. [Google Scholar] [CrossRef] [PubMed]
  41. Martins, L.M.D.R.S.; Pombeiro, A.J.L. Carbon-scorpionate Complexes in Oxidation Catalysis. In Advances in Organometallic Chemistry and Catalysis, the Silver/Gold Jubilee ICOMC Celebratory Book; Pombeiro, A.J.L., Ed.; John Wiley & Sons: Hoboken, NJ, USA, 2014; Chapter 22; pp. 285–294. [Google Scholar]
  42. Milunovic, M.N.M.; Martins, L.M.D.R.S.; Alegria, E.C.B.A.; Pombeiro, A.J.L.; Krachler, R.; Trettenhahn, G.; Turta, C.; Shova, S.; Arion, V.B. Hexanuclear and Undecanuclear Iron(III) Carboxylates as Catalyst Precursors for Cyclohexane Oxidation. Dalton Trans. 2013, 42, 14388–14401. [Google Scholar] [CrossRef] [PubMed]
  43. Sutradhar, M.; Kirilova, 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] [PubMed]
  44. 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] [PubMed]
  45. Silva, T.F.S.; Alegria, E.C.B.A.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Scorpionate vanadium, iron and copper complexes as selective catalysts for the peroxidative oxidation of cyclohexane under mild conditions. Adv. Synth. Catal. 2008, 350, 706–716. [Google Scholar] [CrossRef]
  46. Contaldi, S.; di Nicola, C.; Garau, F.; Karabach, Y.Y.; Martins, L.M.D.R.S.; Monari, M.; Pandolfo, L.; Pettinari, C.; Pombeiro, A.J.L. New coordination polymers based on triangular [Cu3(m3-OH)(m-pz)3]2+ units and unsaturated carboxylates. Molecular structures, electrochemical behaviour and catalytic peroxidative oxidation of cycloalkanes. Dalton Trans. 2009, 4928–4941. [Google Scholar] [CrossRef] [PubMed]
  47. Silva, T.F.S.; Mishra, G.S.; Silva, M.F.G.; Wanke, R.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. CuII complexes bearing the 2,2,2-tris(1-pyrazolyl)ethanol or 2,2,2-tris(1-pyrazolyl)ethyl methanesulfonate scorpionates. X-ray structural characterization and application in the mild catalytic peroxidative oxidation of cyclohexane. Dalton Trans. 2009, 9207–9215. [Google Scholar] [CrossRef] [PubMed]
  48. Nesterov, D.S.; Chygorin, E.N.; Kokozay, V.N.; Bon, V.V.; Boča, R.; Kozlov, Y.N.; Shul’pina, L.S.; Jezierska, J.; Ozarowski, A.; Pombeiro, A.J.L.; et al. Heterometallic CoIII4FeIII2 Schiff Base Complex: Structure, Electron Paramagnetic Resonance, and Alkane Oxidation Catalytic Activity. Inorg. Chem. 2012, 51, 9110–9122. [Google Scholar] [CrossRef] [PubMed]
  49. Silva, T.F.S.; Luzyanin, K.V.; Kirilova, M.V.; Guedes da Silva, M.F.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Novel scorpionate and pyrazole dioxovanadium complexes, catalysts for carboxylation and peroxidative oxidation of alkanes. Adv. Synth. Catal. 2010, 352, 171–187. [Google Scholar] [CrossRef]
  50. Silva, T.F.S.; Guedes da Silva, M.F.; Mishra, G.S.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Synthesis and structural characterization of iron complexes with 2,2,2-tris(1-pyrazolyl)ethanol ligands: application in the peroxidative oxidation of cyclohexane under mild conditions. J. Organomet. Chem. 2011, 696, 1310–1318. [Google Scholar] [CrossRef]
  51. Mishra, G.S.; Silva, T.F.S.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Scorpionate V(III-V) Complexes as Catalyst Precursors for Solvent-free Cyclohexane Oxidation with Dioxygen. Pure Appl. Chem. 2009, 81, 1217–1227. [Google Scholar] [CrossRef]
  52. Silva, T.F.S.; MacLeod, T.C.O.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Schiavon, M.A.; Pombeiro, A.J.L. Pyrazole or tris(pyrazolyl)ethanol oxo-vanadium(IV) complexes as homogeneous or supported catalysts for oxidation of cyclohexane under mild conditions. J. Mol. Catal. A Chem. 2013, 367, 52–60. [Google Scholar] [CrossRef]
  53. Martins, L.M.D.R.S.; Peixoto de Almeida, M.; Carabineiro, S.A.C.; Figueiredo, J.L.; Pombeiro, A.J.L. Heterogenisation of a C-scorpionate Fe(II) complex in carbon materials for cyclohexane oxidation with hydrogen peroxide. ChemCatChem 2013, 5, 3847–3856. [Google Scholar] [CrossRef]
  54. Martins, L.M.D.R.S.; Martins, A.; Alegria, E.C.B.A.; Carvalho, A.P.; Pombeiro, A.J.L. Gold Nanoparticles Supported on Carbon Materials for Cyclohexane Oxidation with Hydrogen Peroxide. Appl. Cat. A Gen. 2013, 464–465, 43–50. [Google Scholar] [CrossRef]
  55. Mishra, G.S.; Alegria, E.C.B.; Martins, L.M.D.R.S.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L. Cyclohexane oxidation with dioxygen catalyzed by supported pyrazole rhenium complexes. J. Mol. Catal. A Chem. 2008, 285, 92–100. [Google Scholar] [CrossRef]
  56. Peixoto de Almeida, M.; Martins, L.M.D.R.S.; Carabineiro, S.A.C.; Lauterbach, T.; Rominger, F.; Hashmi, A.S.K.; Pombeiro, A.J.L.; Figueiredo, J.L. Homogeneous and Heterogenised New Gold C-Scorpionate Complexes as Catalysts for Cyclohexane Oxidation. Catal. Sci. Technol. 2013, 3, 3056–3069. [Google Scholar] [CrossRef]
  57. Kirillov, A.M.; Shul’pin, G.B. Pyrazinecarboxylic acid and analogs: Highly efficient co-catalysts in the metal-complex-catalyzed oxidation of organic compounds. Coord. Chem. Rev. 2013, 257, 732–754. [Google Scholar] [CrossRef]
  58. Kuznetsov, M.L.; Pombeiro, A.J.L. Radical Formation in the [MeReO3]-Catalyzed Aqueous Peroxidative Oxidation of Alkanes: A Theoretical Mechanistic Study. Inorg. Chem. 2009, 48, 307–318. [Google Scholar] [CrossRef] [PubMed]
  59. Kirilova, M.V.; Kuznetsov, M.L.; Romakh, V.B.; Shul’pina, L.S.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L.; Shul’pin, G.B. Mechanism of H2O2 Oxidations Catalyzed by Vanadate Anion or Oxovanadium(V) triethanolaminate (Vanadatrane) in Combination with Pyrazine-2-carboxylic acid (PCA): Kinetic and DFT studies. J. Catal. 2009, 267, 140–157. [Google Scholar] [CrossRef]
  60. Kirilova, M.V.; Kuznetsov, M.L.; Kozlov, Y.N.; Shul’pina, L.S.; Kitaygorodskiy, A.; Pombeiro, A.J.L.; Shul’pin, G.B. Participation of Oligovanadates in Alkane Oxidation with H2O2 Catalyzed by Vanadate Anion in Acidified Acetonitrile: Kinetic and DFT Studies. ACS Catal. 2011, 1, 1511–1520. [Google Scholar] [CrossRef]
  61. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A 2008, A64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  62. Program for Crystal Structure Solution and Refinement; University of Goettingen: Goettingen, Germany, 1997.
  • Sample Availability: Samples of the compounds 13 are available from the authors.

Share and Cite

MDPI and ACS Style

Martins, L.; Nasani, R.; Saha, M.; Mobin, S.; Mukhopadhyay, S.; Pombeiro, A. Greener Selective Cycloalkane Oxidations with Hydrogen Peroxide Catalyzed by Copper-5-(4-pyridyl)tetrazolate Metal-Organic Frameworks. Molecules 2015, 20, 19203-19220. https://doi.org/10.3390/molecules201019203

AMA Style

Martins L, Nasani R, Saha M, Mobin S, Mukhopadhyay S, Pombeiro A. Greener Selective Cycloalkane Oxidations with Hydrogen Peroxide Catalyzed by Copper-5-(4-pyridyl)tetrazolate Metal-Organic Frameworks. Molecules. 2015; 20(10):19203-19220. https://doi.org/10.3390/molecules201019203

Chicago/Turabian Style

Martins, Luísa, Rajendar Nasani, Manideepa Saha, Shaikh Mobin, Suman Mukhopadhyay, and Armando Pombeiro. 2015. "Greener Selective Cycloalkane Oxidations with Hydrogen Peroxide Catalyzed by Copper-5-(4-pyridyl)tetrazolate Metal-Organic Frameworks" Molecules 20, no. 10: 19203-19220. https://doi.org/10.3390/molecules201019203

Article Metrics

Back to TopTop