Euaggelia Skliria,
Ioannis N. Lykakisb and
Gerasimos S. Armatas*a
aDepartment of Materials Science and Technology, University of Crete, Heraklion 71003, Crete, Greece. E-mail: garmatas@materials.uoc.gr
bDepartment of Chemistry, Aristotle University of Thessaloniki, University Campus 54124, Thessaloniki, Greece
First published on 11th September 2014
Multicomponent mesoporous metal oxides show promise in the area of heterogeneous catalysis due to the synergetic interactions between the framework components and the high internal surface area. In this study, we present the synthesis of ordered mesoporous tungsten(VI) oxide–vanadium oxide (V2O5) nanocomposite frameworks via a two-step wet chemical deposition and nanocasting process and demonstrate that they exhibit high catalytic activity and stability for the oxidation of aryl alcohols, using tert-butyl hydroperoxide (t-BuOOH) as oxidant. X-ray diffraction, transmission electron microscopy and nitrogen porosimetry results indicate that the template-free materials possess a 3D mesoscopic structure of discernible domains of parallel-arranged nanorods and have an internal pore surface with narrow mesopores. The chemical composition and molecular structure of the mesoporous matrix were determined with elemental X-ray microanalysis (EDS), diffuse reflectance ultraviolet-visible (UV-vis) and Raman spectroscopy. Our catalytic results indicate that a small addition of V2O5 into the lattice of WO3 has a beneficial effect on the catalytic performance. Thus, the 4% V2O5-loaded WO3 catalyst shows a large improvement in the oxidation of various para-substituted aryl alcohols with respect to the pure mesoporous WO3, giving good-to-high yields (ca. 80–100%) of the target products within 1–4 h reaction time.
Recently, the templated synthesis via the nanocasting route has achieved ambiguous success in producing well-ordered mesoporous materials with controllable composition and textural properties.6 In general, key steps in this synthetic process involve infiltration of suitable metal precursors within the nanopores of the solid template, thermal decomposition at elevated temperature, and removal of the host matrix by selective etching in aqueous NaOH or HF solution. The resulting mesoporous solids, different from those prepared by conventional sol–gel and co-precipitation routes, possess three-dimensional (3D) nanoscale porous structure with regular size and shape imposed by the template pore morphology. In the last few years, a diverse range of ordered mesoporous metal-oxides, such as Co3O4,7 Cr2O3,8 Fe2O3,9 WO3,10 CuO,11 NiO12 and MnO2,13 and mixed metal-oxides, such as CuFe2O414 and Cu/CeO2,14,15 with high crystallinity and large surface area have been successfully prepared by using nanocasting method. More recently, we used hard-templating of mesoporous silica to nanocast well-ordered mesostructured frameworks consisting of nanocrystalline metal oxides (e.g., Co3O4 and Cr2O3) and Keggin-type polyoxometalate clusters (e.g., H3PW12O40 and H3PMo12O40).16 These mesoporous composite materials have a 3D open-pore structure and show great promise in catalytic organic reactions.
In this study, we present the synthesis and catalytic properties of ordered mesoporous frameworks composed of tungsten(VI) oxide and vanadium oxide (V2O5) compounds. As an n-type semiconductor with strong acid activity, WO3 has been successfully used in many catalytic reactions such as hydrodesulfurization of thiophene,17 isomerization of alkanes18 and metathesis of olefins.19 The acid properties of WO3 have also been employed to improve the selective reduction of NOx.20 On the other hand, vanadium–tungsten mixed oxides are highly efficient catalysts, widely applied for the selective reduction of NOx by ammonia21 and oxidation of ethanol22 and volatile organic compounds (VOCs)23 including polychlorinated dibenzofurans (PCDFs).24 Here we demonstrate that mesoporous binary WO3/V2O5 compounds, prepared by a hard template-assisted route, are highly effective and stable catalysts for the oxidation of aromatic alcohols, giving the corresponding carbonyl products in excellent conversion yields (80–100% in 1-4 h). Indeed, the specific mesoporous 4% V2O5-loaded WO3 catalyst we report shows superior activity compared to non-templated material of the same composition, and pure mesoporous WO3.
![]() | ||
Scheme 1 Schematic representation for the synthesis of ordered mesoporous x% V2O5/WO3 composite materials. |
Elemental analysis from energy dispersive X-ray spectroscopy (EDS) on x% V2O5/WO3 products indicated an average atomic ratio of W/V that is consistent with a ∼1, ∼4 and ∼6 wt% of V2O5 loading (x), see Table 1. Note that the EDS vanadium contents are consistently lower than those expected from the stoichiometry of reactions probably due to the insufficient infiltration of NH4VO3 compounds into the silica template. However, the present synthetic method is stunningly reproducible and yields mesoporous heterostructures with consistent composition; it was repeated several times giving materials with V2O5 loading with less than 15% deviation, according to the EDS results. Notably, the EDS spectra did not show any signal from silicon, confirming the complete elimination of the silica matrix (see ESI, Fig. S1†).
Sample | Atomic ratioa (W![]() ![]() |
V2O5 loading (wt%) | Surface area (m2 g−1) | Pore volume (cm3 g−1) | Pore size (nm) | Unit cell (nm) | WTb (nm) |
---|---|---|---|---|---|---|---|
a Based on the EDS analysis.b The framework wall thickness is given by WT = a0 − Dp, where ao is the unit cell size and Dp is the diameter of mesopores. | |||||||
meso-WO3 | 22 | 0.03 | 4.0, 11.4 | 10.2 | 6.2 | ||
1% V2O5/WO3 | 97.5![]() ![]() |
1.0 | 23 | 0.04 | 4.1, 10.7 | 10.3 | 6.2 |
4% V2O5/WO3 | 90.9![]() ![]() |
3.8 | 27 | 0.05 | 4.5, 10.8 | 10.3 | 5.8 |
6% V2O5/WO3 | 86.5![]() ![]() |
5.8 | 37 | 0.06 | 4.4, 10.4 | 10.1 | 5.7 |
The mesoporous structure of the templated materials was investigated with low-angle X-ray diffraction (XRD) and transmission electron microscopy (TEM). As indicated by powder XRD (Fig. 1a), the mesoporous tungsten(VI) oxide (meso-WO3) and x% V2O5/WO3 composite solids exhibited a weak but distinct diffraction peak in low-angle range 2θ ∼ 1°, which according to TEM can be assigned to the (100) reflection of hexagonal P6mm structure. The observation of this diffraction peak clearly suggests a mesoscopic order in these materials, although some deformation of the hexagonal array can be considered due to the low intensity of the (100) diffraction, especially in the 4–6% V2O5-loaded WO3 samples. On the basis of the hexagonal symmetry and XRD data, it is possible to calculate the lattice constant (ao) of the pore structure by using the equation a0 = (2/√3)d100, where d100 is the d-spacing of (100) reflection, and the results are shown in Table 1. These values are almost equal to that of the silica template (ca. 10.7 nm, see ESI, Fig. S2†), suggesting good replication of the silica mesostructure.
Fig. 1b presents the wide-angle XRD patterns of the mesoporous meso-WO3 and x% V2O5/WO3 composite samples. It is revealed that the mesoporous products are highly crystallized and exhibit WO3 monoclinic structure; all XRD patterns can be indexed to a monoclinic cell with lattice constants a = 7.297 Å, b = 7.539 Å, c = 7.688 Å and β = 90.91° (JCPDF card no. 43-1035). The structural assignment based on XRD is also collaborated by TEM and Raman spectroscopy experiments (see below). The XRD patterns did not show any peak due to the crystalline phase of V2O5, implying that vanadium oxide species are uniformly distributed over the WO3 matrix, although the existence of very small grain size of vanadium oxide particles cannot be excluded.
Typical TEM images of the mesoporous 4% V2O5/WO3, in Fig. 2a and b, reveal large domains of parallel arrangement of uniform nanorods, in consistent with the [110] direction of the hexagonal structure. By means of this technique, the nanorods diameter is shown to be ∼7 nm that is reasonably comparable to the mesopore size of the silica template, ∼7.4 nm (Fig. S3†). For the investigation of the crystal structure of the 4% V2O5/WO3 sample, high-resolution TEM (HRTEM) images and selected-area electron diffraction (SAED) pattern were obtained. HRTEM image taken from a local area of the framework clearly shows well-resolved lattice fringes throughout the nanorods (Fig. 2c) with a d-spacing of 3.8 Å, which is in accordance with the (020)-spacing of monoclinic WO3. The image also demonstrated that the nanorods are interconnected to each other with short bridges to form mesostructured superlattices. Fig. 2d depicts the SAED pattern of the 4% V2O5/WO3 and shows a series of spotted Debye–Scherrer diffraction rings, suggesting randomly oriented nanocrystals. All these diffraction rings can be readily assigned to the monoclinic phase of WO3, in agreement with XRD results.
Fig. 3 displays the N2 adsorption–desorption isotherms and the corresponding nonlocal density functional theory (NLDFT) plots for mesoporous meso-WO3 and x% V2O5/WO3. All isotherms show type-IV curves with an H3-type hysteresis loop, according to the IUPAC classification, which are attributed to the mesoporous solids with interconnected porosity. In general, the presence of H3 hysteresis in relative pressure (P/Po) range 0.4–0.85 is related to the slit-shaped mesopores.28 The adsorption isotherms also exhibit a weak but distinguishable capillary condensation step at relative pressure (P/Po) ∼ 0.2–0.3, indicative of narrow distribution of pore sizes.29 The mesoporous x% V2O5/WO3 composites have Brunauer–Emmett–Teller (BET) surface areas in the range of 23–37 m2 g−1 and total pore volumes in the range of 0.04–0.06 cm3 g−1. The mesoporous meso-WO3 show a surface area of 22 m2 g−1 and a total pore volume of 0.03 cm3 g−1, which are slightly lower than those of composite materials possibly due to the heavier structure of WO3 (7.2 g cm−3) relative to V2O5 (3.3 g cm−3).
The pore width in as-prepared materials was determined by using the pore size analysis of NLDFT adsorption model for slit-shaped pores, and was found to be ∼4–5 nm (insets of Fig. 3). This pore size reflects the void space between the interconnected nanorods, which is very close to the framework wall thickness of the SBA-15 template (ca. 3.3 nm, see ESI, Fig. S3†). The broad shoulder at 10–11 nm associated the pore size distributions is corresponding to the large voids between the partially interconnected nanorods. From a combination of data from NLDFT and XRD analysis, the pore wall thickness is calculated to be about 6–7 nm, in agreement with TEM analysis. These results give evidence that the mesoporous products are good replicas of the silica template. Table 1 summarizes the analytical data and the morphological properties of mesoporous meso-WO3 and x% V2O5/WO3 composite materials.
The molecular structure of WO3 and V2O5 components in mesoporous matrix was investigated with diffuse reflectance ultraviolet-visible (UV-vis) and Raman spectroscopy. The UV-vis spectra of the as-prepared samples, transformed from the diffuse reflection data according to the Kubelka–Munk theory,30 show a sharp optical absorption onset in the energy range ∼430–450 nm (∼2.8–2.9 eV), which is interpreted by the interband electron transitions in WO3 (see Fig. S4 of the ESI†). The broad absorption band in the region between 550 and 650 nm appeared in the UV-vis spectra of x% V2O5/WO3 composites is assigned to the d–d charge transitions of V2O5 species.31
Raman spectroscopy is a very efficient technique to probe the crystal structure of WO3 materials.32 The Raman spectra of meso-WO3 and x% V2O5/WO3 materials, shown in Fig. 4, are consistent with the monoclinic phase of WO3. All spectra displayed intense peaks in 793–802 and 679–709 cm−1 regions that correspond to the W–O–W stretching mode and broad peaks at 324 and 265–273 cm−1 due to the bending modes of O–W–O bonds in the monoclinic WO3, respectively.32,33 Compared to the Raman spectrum of meso-WO3, the W–O–W stretching bands in composite samples shift toward lower wavenumbers, possibly due to the additional formation of W–O–V bonds. The shifts from 120 to 130 cm−1 are attributed to the lattice vibration of crystalline WO3.34 Clear evidence for the inclusion of V2O5 compounds in mesoporous structure comes from the Raman shift in the region between 988 and 995 cm−1. This peak corresponds to the stretching mode of VO bonds in the crystalline V2O5.35 Prominent blue shift of this band, especially in 6% V2O5 loaded sample, may be related with the presence of polymeric vanadia species in the mesoporous structure.36 Taken together with high-resolution TEM images, these results suggest that the pore walls are composed of WO3 nanocrystals and a small quantity of V2O5 compounds.
![]() | ||
Fig. 4 Raman spectra of mesoporous (a) meso-WO3 and (b) 1% V2O5/WO3, (c) 4% V2O5/WO3 and (d) 6% V2O5/WO3 composite materials. |
Having established the optimal reaction conditions for the oxidation of 1, the compositional dependence of x% V2O5/WO3 mesoporous on catalytic activity was studied. Catalytic results shown in Fig. 5 and Table 2 indicated that V2O5 compounds included into the WO3 matrix have an appealing effect on the catalytic performance. In particular, WO3 samples loaded with 4–6 wt% V2O5 afford a moderate-to-high yield of 1a (63–89% in 2 h), while the composite material containing lower amount of V2O5 (∼1 wt%) shows a noticeable drop in activity (55% yield of 1a). Of particular note, meso-WO3 shows little catalytic activity under the same conditions (ca. 15% conv. of 1, in 2 h). On the basis of 1 consumption, the mesoporous 4% V2O5/WO3 was found to be the best catalyst under the present conditions, giving 89% yield of 1a in 2 h. Indeed, this catalyst, unlike the other examined materials, catalyzed the oxidation of 1 almost quantitatively to ketone 1a (∼97% yield) within 3 h (see Table 2). Control experiments did not show any significant catalytic activity in the absence of catalyst; the conversion of 1 was less than 3% after 2 h reaction.
![]() | ||
Fig. 5 Time-dependent conversion plots for the oxidation of 1-phenylethanol (1) by mesoporous meso-WO3 and x% V2O5/WO3 composite materials and macroscopic bulk-4% V2O5/WO3 solid. |
Catalyst | Conversionb (%) | Selectivity (%) | Kinetic constant, k (min−1) |
---|---|---|---|
a Experimental conditions: 0.1 mmol 1-phenylethanol, 50 mg catalyst, 40 equiv. t-BuOOH, 2 mL CH3CN, 50 °C, 2 h.b Determined by GC-MS analysis. | |||
meso-WO3 | 15 | 100 | 0.004 |
1% V2O5/WO3 | 55 | 100 | 0.010 |
4% V2O5/WO3 | 89 | 100 | 0.033 |
6% V2O5/WO3 | 63 | 100 | 0.041 |
S-4% V2O5/WO3 | 84 | 100 | 0.021 |
Bulk-4% V2O5/WO3 | 80 | 100 | 0.022 |
The time evolution of the 1 oxidation can be sufficiently described by a pseudo-first-order reaction model. This is reasonable if we account the excess of t-BuOOH oxidant, so that its concentration could be considered constant during the reaction. Fig. S7 (ESI†) shows the plots of ln(Ct/Co) versus time (where Co and Ct are the concentrations of 1 at the initial state of reaction and at the time t, respectively), by which the apparent first-order reaction rates (k) were obtained as a slope of the linear fits. The corresponding k values are shown in Table 2. Kinetic analysis indicated that the oxidation reaction proceeds much faster over 4% V2O5/WO3 (0.033 min−1) than the meso-WO3 (0.004 min−1) and 1% V2O5/WO3 (0.010 min−1) catalysts. The 6% V2O5/WO3 although oxidizes 1 at a faster rate (0.042 min−1), it gives moderate conversion yield (∼66%) of 1a. These results indicate that addition of small amount of V2O5 into the WO3 lattice has a beneficial effect of improving the catalytic activity of WO3. It seems that vanadium oxide is really synergistic catalyst, where WO3 component is activated by the V2O5 species that solid-dissolved in mesoporous matrix. In agreement with this assumption, the W–O–V contribution in composite catalysts is collaborated by Raman spectroscopy.
To examine the role of W–O–V sites on the catalytic performance of V2O5/WO3 materials, we also prepared mesoporous 4% V2O5-loaded WO3 catalyst featuring a V2O5-poor surface composition, and then we examined its catalytic activity under similar conditions. This catalyst, designated as S-4% V2O5/WO3, was synthesized by following a procedure similar to that for 4% V2O5/WO3, but using SBA-15 silica as template. The constitution of the non-functionalized silica template used in this experiment is expected to produce mesostructured V2O5/WO3 nanorods in which a certain amount of V2O5 will be located to the internal structure. Therefore, this catalyst will possess less V2O5 active species accessible to reactants than its modified silica-templated 4% V2O5/WO3 counterpart. Remarkably, the S-4% V2O5/WO3 exhibited lower activity than the corresponding mesoporous 4% V2O5/WO3 sample in oxidation of 1, giving 84% conversion yield of 1a in 2 h with a reaction rate constant of 0.021 min−1, see Table 2. This reflects that W–O–V sites on the surface of x% V2O5/WO3 eventually contribute to the high catalytic efficiency. Notably, the mesoporous 4% V2O5/WO3 achieves also higher oxidation kinetic as compared to its non-porous analog. For purpose of comparison, we also performed the oxidation experiment on non-templated 4% V2O5/WO3 composite solid, denoted as bulk-4% V2O5/WO3, which is prepared by solid phase sintering of a powder blend containing NH4VO3 and 12-phosphotungtic acid compounds; the product shows a BET surface area of 9 m2 g−1 and a pore volume less than 0.01 cm3 g−1. Remarkably, the bulk-4% V2O5/WO3 microparticles although afforded excellent yield of 1a (∼94%) in 4 h, results to less efficient reaction rate (0.022 min−1) that does mesoporous 4% V2O5/WO3 under similar conditions (see ESI, Fig. S7†). Such superiority of the mesoporous 4% V2O5/WO3 material may be related to the solid solution of V2O5 oxides into the WO3 lattice, the high crystallinity of WO3 and the three-dimensional open-pore structure, which offer competitive advantages to the activation of WO3 structure.
To test the recycling ability of our catalyst, we carried out repeated oxidations of 1 using 4% V2O5/WO3 as catalyst. After each reaction, the catalyst was recovered by simple filtration, washed several times with acetonitrile, and then reused for the next catalytic run. As shown in Fig. 6, the acetophenone (1a) yield remained essentially constant (∼94–98%) even after four consecutive catalytic cycles, reflecting high durability and reusability of the catalyst. The stability of the mesoporous structure was verified by elemental X-ray microanalysis and N2 physisorption measurements. EDS spectra indicated no detectable leaching of V2O5 after catalysis, showing an average W/V atomic ratio that corresponds to a V2O5 content of about 3.6 wt%. Nitrogen adsorption data evidenced no change in the mesoporous structure of reused catalyst compared to the fresh material (ESI, Fig. S8†), indicating a surface area of 26 m2 g−1 and pore volume of 0.05 cm3 g−1. These results consist with high stability and reusability of the 4% V2O5/WO3 mesostructure. In addition, the same catalytic reaction was also conducted by using 4% V2O5/WO3 catalyst. When the catalyst was separated from the reaction mixture shortly (30 min) and the reaction filtrate was further stirred at 50 °C, no additional conversion of 1 was detected by GC-MS analysis even after 2 h; we obtained a ∼58% and ∼59% conversion yield of 1 before and after removal of the catalyst. These results provide strong evidence that the present oxidation reactions are heterogeneous in nature.
The mesoporous 4% V2O5/WO3 efficiently catalyzes the oxidation of various para-substituted aromatic alcohols, such as 1-phenylethanols and benzyl alcohols, to their target products. As seen in Table 3 and Fig. S9a of the ESI,† all the substituted 1-phenylethanols (2–6) were oxidized into the corresponding ketones at an almost quantitatively yield (>97%) within 3 h. Also, oxidation of substituted benzyl alcohols (7–9) afforded the corresponding aryl aldehydes in a range of 80–94% yields, although in prolonger reaction time (4 h). Of particular note, p-methyl benzyl alcohol (8) was oxidized into the corresponding p-methyl benzaldehyde (8a) as the major product (55% conv., 80% relative yield) in 15 min, while at the longer reaction time (3 h) the carboxylic acid (p-methylbenzoic acid, 8b) was formed as the only product. Evidence for this was obtained from GC-MS and NMR spectroscopy. To rule out the possibility of the self-oxidation reaction of aldehyde 8a to carboxylic acid 8b, we reacted the aldehyde 8a with t-BuOOH but in the absence of catalyst. In this experiment, only a 8% conversion yield of the p-methylbenzoic acid (8b) was observed in the reaction mixture after 2 h, demonstrating that oxidation of aromatic aldehyde to the corresponding carboxylic acid is a catalytic process. These results suggest the strong oxidizing character of the 4% V2O5/WO3 material.
Substrate | Product | Yieldb (%)/time (h) | Kinetic const., k (min−1) | |
---|---|---|---|---|
a Experimental conditions: 0.1 mmol substrate, 50 mg catalyst, 40 equiv. t-BuOOH, 2 mL CH3CN, 50 °C.b Determined by GC-MS, with error ±1%. | ||||
1 | ![]() |
![]() |
97/3 | 0.033 |
2 | ![]() |
![]() |
100/1 | 0.150 |
3 | ![]() |
![]() |
99/3 | 0.057 |
4 | ![]() |
![]() |
98/3 | 0.051 |
5 | ![]() |
![]() |
97/3 | 0.052 |
6 | ![]() |
![]() |
96/3 | 0.043 |
7 | ![]() |
![]() |
80/4 | 0.025 |
8 | ![]() |
![]() |
100/3 | 0.056 |
9 | ![]() |
![]() |
94/4 | 0.026 |
10 | ![]() |
![]() |
99/2 | 0.041 |
It is noteworthy that the presence of electron-donating or electron-withdrawing group has a moderate effect on the catalytic activation of aromatic alcohols. For example, the electron rich alcohols 2 (X = –OCH3) reacted to form the corresponding ketone in excellent (>99%) yield in 1 h reaction time. Similarly, alcohols bearing electron-withdrawing substituent such as p-bromobenzyl alcohol (4), p-chlorobenzyl alcohol (5) and p-nitrobenzyl alcohol (6) were oxidized to the corresponding carbonyl compounds with 96–98% conversion. However, the oxidation reaction proceeds slightly faster as the electron-donating ability of the substituent functionality increases, see Fig. S9b of the ESI.† Specifically the pseudo first-order reaction rates, shown in Table 3, indicate an about three times faster kinetic rate for 2 (MeO-substituted) oxidation relative to the oxidation of 3 (Me-substituted), 5 (Cl-substituted) and 6 (NO2-substituted); kMeO/kMe = 2.6, kMeO/kCl = 2.9 and kMeO/kNO2 = 3.5. Similar, p-methylbenzyl alcohol (8) was also oxidized to the corresponding aldehyde (8a) in a faster reaction rate (approximately two times) compared to the electron poor alcohols 7 and 9, i.e. possessing the electron-deficient H- and NO2-groups in para position, respectively. It should be stressed that steric properties of the substituent seem to not affect significantly the reaction process. For example, in the case of substrate 10 where the α-substituent is a phenyl group, the rate of the catalytic reaction is about 1.3-times higher to that of 2 alcohol, which contained a methyl group next to the benzylic carbon. These results clearly show that our 4% V2O5/WO3 catalyst is able to catalyze the oxidation of hindered primary and secondary aromatic alcohols with high efficiencies in presence of t-BuOOH.
Footnote |
† Electronic supplementary information (ESI) available: EDS spectra for x% V2O5/WO3 composites, XRD and N2 physisorption data for SBA-15 and APS/SBA-15 templates, diffuse reflectance UV-vis spectra and catalytic data for meso-WO3 and x% V2O5/WO3, and catalytic data and N2 adsorption–desorption isotherms of reused 4% V2O5/WO3 catalyst. See DOI: 10.1039/c4ra07850k |
This journal is © The Royal Society of Chemistry 2014 |