Zeinab Asgharpoura,
Faezeh Farzaneh*a and
Alireza Abbasib
aDepartment of Chemistry, Faculty of Physics & Chemistry, Alzahra University, P.O. Box 1993891176, Vanak, Tehran, Iran. E-mail: faezeh_farzaneh@yahoo.com; Farzaneh@alzahra.ac.ir; Fax: +98-21-88041344; Tel: +98-21-88258977
bSchool of Chemistry, College of Science, University of Tehran, P.O. Box 14155 6455, Tehran, Iran
First published on 3rd October 2016
The [Co(2-hydroxy-1-naphthaldehyde)2(DMF)2] complex 1 was prepared and characterized by X-ray crystallography, successfully supported on modified Fe3O4 nanoparticles using tetraethylorthosilicate (TEOS) and (3-aminopropyl)trimethoxysilane (APTMS) and designated as Fe3O4@SiO2@APTMS@complex nanocatalyst. FT-IR, elemental analysis, XRD, VSM and TEM studies have been used to characterize the nanocatalyst. The catalytic activity of Fe3O4@SiO2@APTMS@complex was evaluated by the epoxidation of norbornene, cyclooctene, styrene, α-methyl styrene, trans-stilbene and limonene, with 50–100% conversions and 83–100% selectivities. Furthermore, it was found that Fe3O4@SiO2@APTMS@complex successfully catalyzes the oxidation of activated secondary alkanes such as fluorene, adamantane, diphenyl methane, and ethylbenzene with 40–100% conversions and 80–100% selectivities toward the corresponding products. Based on the obtained results, the heterogeneity and reusability of the catalyst seems promising.
On the other hand, the coordination chemistry of ligands derived from catechols and o-benzoquinones has been developed, primarily, over the past twenty five years with interesting and surprising results.21–23 Initial interest in these complexes was associated with the redox activity of the quinone ligands and the potential for forming compounds that may exist in a number of electronic states due to the combined electrochemical activity of the cobalt ion and one or more quinone ligands.23
Homogeneous catalysis has been practiced for a very long time in large-scale oxidation reactions. However, these homogeneous catalysts face the problem of separation from the reaction mixture and reuse; they very often decompose during the catalytic reactions. A homogeneous complex catalyst can be recovered and reused if it is heterogenized on an insoluble support.24,25 The dispersion and attachment of catalytically-active metallic nanoparticles or organometallic complexes onto polymeric or inorganic solid support materials such as silica,26 inorganic layers,27 microporous and mesoporous such as zeolites, MCM-41 and SBA15,28–30 and magnetic nanoparticles31,32 via non-covalent or covalent interactions can generally produce effective heterogeneous catalysts for numerous reactions including oxidation of alkanes and alkenes.33–36 In this study, a new cobalt complex with 2-hydroxynaphthaldehyde was prepared and characterized followed by immobilization on modified iron oxide nanoparticles as heterogeneous catalyst for oxidation reactions of alkanes and alkenes.
![]() | ||
Fig. 1 An ORTEP view of complex 1 drawn at 50% ellipsoid probability levels with the crystallographic numbering. |
a Symmetry code: (i) −x + 1, −y + 1, −z. | |
---|---|
Bond lengths (Å) | |
Co(1)–O(2)i | 1.989 (2) |
Co(1)–O(2) | 1.989 (2) |
Co(1)–O(3)i | 2.059 (2) |
Co(1)–O(3) | 2.059 (2) |
Co(1)–O(1)i | 2.191 (2) |
Co(1)–O(1) | 2.191 (2) |
![]() |
|
Bond angles (°) | |
O(2)i–Co(1)–O(2) | 180.0 |
O(2)i–Co(1)–O(3)i | 87.71 (8) |
O(2)–Co(1)–O(3)i | 92.29 (8) |
O(2)i–Co(1)–O(3) | 92.29 (8) |
O(2)–Co(1)–O(3) | 87.71 (8) |
O(3)i–Co(1)–O(3) | 180.0 |
O(2)i–Co(1)–O(1)i | 88.97 (9) |
O(2)–Co(1)–O(1)i | 91.03 (9) |
O(2)–Co(1)–O(1) | 88.97 (9) |
O(3)i–Co(1)–O(1) | 90.02 (9) |
O(3)–Co(1)–O(1) | 89.71 (9) |
O(1)i–Co(1)–O(1) | 180.000 (1) |
The crystal structure of 1 was satisfactorily described in the space group P with Co(II) ion in a distorted octahedral 6-coordinated geometry. Bond angles in the octahedron vary from 88.97(9)° to 91.03(9)° for cis and 180° for trans-substituted, indicating the small distortion degree in the octahedron.
The asymmetric unit of 1 is composed of one DMF (dimethylformamide) molecule and one 2-hydroxy-1-naphthaldehyde anion. Two 2-hydroxy-1-naphthaldehyde ligands coordinate to Co(II) as bidentate [O2, O3], chelating ligand via the deprotonated carboxylic acid and phenoxy oxygen atoms, defining the equatorial coordination plane. There is a distinct trend in Co–O bond lengths in the complex with Co1–O2 (carboxyl) < Co1–O3 (phenoxyl) < Co1–O1 (dimethylformamide). The complexes are held together through hydrogen bonds (C3–H3A⋯O3) (C3–H3A: 0.9600 (Å), H3A⋯O3: 2.4800(Å), C3⋯O3: 3.353(4) (Å), C3–H3A⋯O3: 151.00°) making a chain along [100] (Fig. 2). The neighboring chains are then connected by pi⋯pi interactions (Cg(3)⋯Cg(4), 4.161 Å, where Cg(3): C4/C5/C7/C8/C9/C10 and Cg(4): C7/C8/C11/C12/C13/C14) stabilizing 3D framework (Fig. 3).
The FT-IR spectra of free 2-hydroxy-1-naphthaldehyde as ligand and [Co(2-hydroxy-1-naphthaldehyde)2(DMF)2] complex 1 are shown in Fig. 4a and b, respectively. The bands appearing at 3100, 2855, 1647, 1180 and 770 cm−1 are due to the phenolic OH, C–H, aldehyde CO, phenolic C–O stretching and O–H bending, respectively (Fig. 4a). The absence of absorption bands at 3100 and 770 cm−1 due to the stretching and bending of phenolic OH in the FT-IR spectrum of complex 1 confirms the bond formation between cobalt ion and phenolic oxygen via deprotonation. Moreover, the 10 cm−1 increase in the wave number of band appearing at 1119 cm−1 is due to the phenolic C–O absorption which also confirms the CO coordination to the metal.37 The red shift observed in C
O wave number appearing at 1641 cm−1 indicates the oxygen atom coordination to the cobalt ion with no need to enolization.38
![]() | ||
Fig. 4 FT-IR spectra of (a) 2-hydroxy-1-naphthaldehyde and (b) [Co(2-hydroxy-1-naphthaldehyde)2(DMF)2] complex 1. |
The electronic absorption spectrum of complex 1 is shown in Fig. S1.† The cobalt(II) complex 1 in DMF exhibits three bands at 266, 313 and 369 nm, and a broad band with low absorptivity at 416 nm. The first band observed at 266 nm can be assigned to the aromatic ring transition π → π*. The other bands displaying at 313 and 369 are due to n → π* type electronic transitions. Broad band with low absorptivity centred at 416 nm can be assigned to a d–d transition from the 3d cobalt(II) metal center.39
To evaluate the thermal stability of complex 1, thermal decomposition was studied (Fig. 5). The TGA curve shows two step weight loss at 133 and 348 °C together with exothermic peaks in the DTG curve due to the decomposition of the dimethylformamide and organic ligand (obs.: 89.5%, calc.: 89.2%) with the formation of cobalt oxide.
As shown in Fig. 6a, the broad band with relatively low intensity centered at 3400 and 1600 cm−1 should be attributed either to the surface Fe–OH groups, or the stretching and bending vibrations of the adsorbed water. The stretching vibration of Fe–O also appears at 555 cm−1. The band appearing at 1091 cm−1 due to Si–O stretching demonstrates the presence of SiO2 components (Fig. 6b).40–42 After functionalization of silica coated magnetic nanoparticles with APTMS, new bands displaying at 2870–2920 cm−1 are due to the N–H and C–H stretching vibrations.43 The FT-IR spectrum of catalyst 1 is shown in Fig. 6d. Upon immobilization of cobalt complex on the modified nanoparticles, the appearance of two peaks at 1647 and 1180 cm−1 due to the ν(CO) and ν(C–O), respectively confirms the occurrence of immobilization.
![]() | ||
Fig. 6 The FTIR spectra of (a) Fe3O4, (b) Fe3O4@SiO2, (c) Fe3O4@SiO2@APTMS, (d) complex 1, (e) Fe3O4@SiO2@APTMS@complex 1 before using and (f) after using as catalyst. |
To study the magnetic properties of magnetite nanoparticles, the hysteresis loops of magnetite and functionalized magnetite nanoparticles at room temperature were investigated by using vibrating sample magnetometry (VSM). Magnetization curves of Fe3O4, Fe3O4@SiO2, Fe3O4@SiO2@APTMS, Fe3O4@SiO2@APTMS@complex 1 before and after using as catalyst are shown in Fig. 7a–c, respectively. The magnetization curve of Fe3O4 exhibited superparamagnetic property with saturation magnetization of about 61 emu g−1. The silica coated Fe3O4 magnetite nanoparticles also showed superparamagnetic behavior with decreased saturation magnetization of about 57 emu g−1. After coating of Fe3O4@SiO2 nanoparticles with aminosilane, the superparamagnetic behavior of Fe3O4@SiO2@APTMS was found to be about 56 emu g−1. Immobilization complex 1 within the Fe3O4@SiO2@APTMS surface exhibited superparamagnetic property decrease to saturation magnetization of about 51 emu g−1. Recall that the superparamagnetic property of Fe3O4@SiO2@APTMS@complex 1 prevents aggregation and enables it to disperse rapidly when the magnetic field is removed.40,42,44
![]() | ||
Fig. 7 Magnetization curves of (a) Fe3O4, (b) Fe3O4@SiO2, (c) Fe3O4@SiO2@APTMS, (d) Fe3O4@SiO2@APTMS@complex 1 before and (e) after using as catalyst. |
The X-ray diffraction pattern of (a) Fe3O4, (b) Fe3O4@SiO2, (c) Fe3O4@SiO2@APTMS, (d) complex 1 (e) Fe3O4@SiO2@APTMS@complex 1 before and (f) after using as catalyst are shown in Fig. 8a–f, respectively. As shown in Fig. 8a, it can be seen that the obtained Fe3O4 has highly crystalline cubic spinel structure in agree with the standard Fe3O4 (cubic phase) XRD spectrum (JCPDS no. 19-0629).45,46 The similar set of characteristic peaks observed for Fe3O4@SiO2 and Fe3O4@SiO2@APTMS (Fig. 8b and c) indicate the stability of the crystalline phase of silica coating and surface functionalization. The XRD pattern of complex 1 and Fe3O4@SiO2@APTMS@complex 1 are shown in Fig. 8d and e, respectively. Observation of two additional peak at 2 theta 15 and 18 for Fe3O4@SiO2@APTMS@complex 1 should be attributed to the immobilization of complex 1 on modified iron oxide magnetic nanoparticles.
![]() | ||
Fig. 8 XRD pattern for (a) Fe3O4, (b) Fe3O4@SiO2, (c) Fe3O4@SiO2@APTMS (d) complex 1 (e) Fe3O4@SiO2@APTMS@complex 1 before using and (f) after using as catalyst. |
The EDX of Fe3O4@SiO2@APTMS@complex 1 confirmed the presence of cobalt complex on modified iron nanomagnet (Fig. 9).
The TEM image of Fe3O4@SiO2@APTMS@complex 1 presented in Fig. 10A and B before and after using it as catalyst show the core and shell of the prepared nanoparticles with the average particle size of 10 nm.
![]() | ||
Fig. 11 Conversion and selectivity of norbornene with different amounts of Fe3O4@SiO2@APTMS@complex 1 as catalyst 2. |
![]() | ||
Fig. 12 Conversion and selectivity of norbornene at different times in the presence of the Fe3O4@SiO2@APTMS@complex 1 as catalyst 2. |
![]() | ||
Fig. 13 The effect of solvent on the conversion and selectivity of norbornene using Fe3O4@SiO2@APTMS@complex 1 as catalyst 2. Ethylbenzene were investigated under optimized reaction conditions. |
The nature of solvent has an important role on the conversion and product selectivity. In this study, both protic and aprotic solvents were used to investigate the solvent effect on the norbornene epoxidation. As seen in Fig. 13, protic solvents such as methanol and ethanol were not suitable for norbornene epoxidation. These results are similar to those previously reported for alkene epoxidation studies.47–49 On the other hand, high conversion and selectivity toward epoxide was achieved in aprotic solvents in the order of n-hexane < CHCl3 < CH2Cl2 < CH3CN. The observed trend which is consistent with increasing order of dielectric constants clearly indicates the increasing solvent polarity role toward high conversion (100%) and high epoxide selectivity (100%) in norbornene epoxidation. When epoxidation of norbornene was carried out in the presence of both complex 1 (catalyst 1) and Fe3O4@SiO2@APTMS@complex 1 as catalyst 2, 100% conversion and selectivity was observed within 6 h in CH3CN. Inspired by these results, other alkenes such as cyclooctene, styrene, α-methyl styrene, trans-stilbene and cyclohexene and activated secondary alkanes including fluorene, adamantane, diphenylmethane and (Tables 2 and 3). As seen in Table 2, the reaction results can be categorized in three trends. Whereas norbornene or cyclooctene were selectively converted to the corresponding epoxides (entries 1 and 2, Table 2), epoxidation of α-methyl styrene, styrene and trans-stilbene proceeded to the corresponding epoxides as well as another by-product in each case (entries 3 to 5, Table 2). On the other hand, cyclohexene afforded cyclohexene epoxide plus 2-cyclohexene-1-ol and 2-cyclohexene-1-one (entry 5, Table 2). Interestingly, fluorene, adamantane, diphenylmethane and ethylbenzene were oxidized to the corresponding ketones with 40–100% conversions and 100% selectivities (Table 3).
Entry | Substrate conversiona (%) | Epoxide selectivityb (%) (TON) | ||
---|---|---|---|---|
Catalyst 1 | Catalyst 2 | Catalyst 1 | Catalyst 2 | |
a Reaction conditions: catalyst: 25 mg, substrate: alkenes (10 mmol), (trans-stilbene, 1 mmol), alkanes (1 mmol) (cyclooctane, 10 mmol), TBHP (14 mmol), solvent (acetonitrile, 5 mL), conditions (reflux at 80 °C), time (1: 6 h, 2, 3 h).b TON is the mmol of product to mmol of vanadium present in catalyst.c Acetophenone was identified as byproduct.d Benzaldehyde (10%) and benzoic acid (15%) were identified as byproducts.e Benzaldehyde (7%) and benzoic acid (13%) were identified as byproducts.f Benzaldehyde (10%) and benzoic acid (20%) were identified as byproducts.g Benzaldehyde (3%) was identified as byproduct.h 2-Cyclohexene-1-ol (42%) and 2-cyclohexene-1-one (23%) were identified as byproducts.i 2-Cyclohexene-1-ol (35%) and 2-cyclohexene-1-one (12%) were identified as byproducts. | ||||
1 | ![]() |
![]() |
||
2 | ![]() |
![]() |
||
3 | ![]() |
![]() |
||
4 | ![]() |
![]() |
||
5 | ![]() |
![]() |
||
6 | ![]() |
![]() |
Entry | Substrate conversiona (%) | Product selectivity (%) | ||
---|---|---|---|---|
Catalyst 1 | Catalyst 2 | Epoxideb (TON) | ||
Catalyst 1 | Catalyst 2 | |||
a Reaction conditions: catalyst, 25 mg, substrate (alkanes, 1 mmol), TBHP (1.4 mmol), solvent (acetonitrile, 5 mL, reflux at 80 °C), time (1: 6 h, 2, 3 h).b TON is the mmol of product to mmol of vanadium present in catalyst.c Diphenyl methanol was identified as byproduct. | ||||
1 | ![]() |
![]() |
||
2 | ![]() |
![]() |
||
3 | ![]() |
![]() |
||
4 | ![]() |
![]() |
To cast light onto the reaction mechanism,50–52 the norbornene epoxidation and fluorene or adamantane oxidation reactions were carried out in the presence of diphenylamine as a radical scavenger. Observation of suppression of the reaction yields clearly revealed that these processes have proceeded via a radical mechanism.53–55 Based on the operation of a radical mechanism, one may rationalize the reaction mechanism (Scheme 2). It seems likely that reaction between catalyst and TBHP initially affords a CoIII–peroxy compound. Subsequently, t-BuOO radical is released from CoIII–peroxy adduct and regenerating the catalyst (Scheme 2a).56 In the next step, t-BuOO radical either is added to the alkene double bond and forming t-butoxperoxy I (path 1, Scheme 2a) or abstract an active H from allylic C–H bond and generating cycloalkenyl radical I′ intermediate (path 1′, Scheme 2a). Whereas elimination of t-BuOH from I affords the epoxide product (path 2, Scheme 2a), reaction of radical I′ with molecular O2 gives rise to cycloalkenylperoxide radical II′ (paths 2′, Scheme 2a), which subsequently generates 2-cycloalkenyl-1-ol and 2-cycloalkenyl-1-one via a disproportionation reaction (paths 3′, Scheme 2a). Since aqueous solution of TBHP was used as oxidant, partial hydrolysis of α-methyl styrene, styrene and trans-stilbene epoxides to the relevant 1,2-diols perhaps catalyzed by some acidic free silanol groups present on the support is inevitable (path 3, Scheme 2a).57 Subsequently, the obtained 1,2-diols undergo extremely rapid scission to acetophenone, benzaldehyde and benzaldehyde, respectively (path 4, Scheme 2a).58 Recall that norbornene and cyclooctene neither undergo hydrolysis since their 1,2-diol degradations are inefficient, nor susceptible to allylic site oxidation due to instability of the bridgehead radical intermediate of the former (Scheme 2b)59,60 and orthogonality of the C–H orbital to the CC double bond in the latter (Scheme 2c) respectively.61 Particularly significant is cyclohexene in which either epoxidation or oxidation take place concomitantly on C
C and C–H bonds, respectively. Since radical p and C
C π-orbitals of cyclohexene are closely parallel to each other, it easily undergoes allylic-site oxidation, affording cyclohexenol and cyclohexenone (Scheme 2d).
![]() | ||
Scheme 2 Proposed reaction mechanism for the epoxidation of alkene and oxidation of alkane with the catalyst. |
Finally, transformation mechanism of alkanes having active secondary C–H bonds such as fluorene, adamantane, diphenyl methane, and ethylbenzene to the corresponding ketones under the catalytic effect of catalysts 1 and 2 is suggested (Scheme 2e).62
The stability of the magnetically recovered Fe3O4@SiO2@APTMS@complex 1 as catalyst 2 was confirmed with spectroscopic studies and elemental analysis. The FTIR and XRD spectra of catalyst 2 before and after using as catalyst are similar. Moreover, as seen in VSM spectra after using Fe3O4@SiO2@APTMS@complex 1 as catalyst 2, the saturation magnetization was found to be 48 emu g−1. Therefore, it can be concluded that there is no significant change in its magnetic property and the magnetic core seems to be stable during epoxidation reaction.63 Cobalt and iron content of Fe3O4@SiO2@APTMS@complex 1 determined before and after using it as catalyst are 3.66%, 41.91% and 3.63%, 41.90%, respectively. Observation of no catalytic activity in the filtrate when it was subjected to the epoxidation conditions in the absence of catalyst clearly indicated that the catalyst is truly heterogeneous in nature.
The reusability of the catalyst 2 after recovery by magnetic decantation and washing with solvent was examined in the epoxidation of norbornene in another run by addition of fresh norbornene similar to the initial reaction. As indicated in Fig. 14, a decrease in conversion from 100 to 97% without any changes in selectivity was observed after five runs.
![]() | ||
Fig. 14 The effect of recycling of Fe3O4@SiO2@APTMS@complex as catalyst (2) on epoxidation of norbornene. |
Whereas the catalytic activity of homogeneous catalyst 1 exhibits high activity in comparison to that of the heterogeneous catalyst 2 toward the formation of corresponding products (Tables 2 and 3), the latter with higher TON as indicated in these tables is considerable.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 1403895. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra18154f |
This journal is © The Royal Society of Chemistry 2016 |