Next Article in Journal
Preparation of TiO2-CNT-Ag Ternary Composite Film with Enhanced Photocatalytic Activity via Plasma-Enhanced Chemical Vapor Deposition
Previous Article in Journal
Combined Biological and Photocatalytic Degradation of Dibutyl Phthalate in a Simulated Wastewater Treatment Plant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tetrabutyl Ammonium Salts of Keggin-Type Vanadium-Substituted Phosphomolybdates and Phosphotungstates for Selective Aerobic Catalytic Oxidation of Benzyl Alcohol

1
Facultad de Ciencias Químicas, Universidad de Concepción, Concepción 4070371, Chile
2
ANID—Millennium Science Initiative Program-Millennium Nuclei on Catalytic Process towards Sustainable Chemistry (CSC), Santiago 7820436, Chile
3
Centro de Investigación y Desarrollo en Ciencias Aplicadas Dr. Jorge J. Ronco, Universidad de La Plata, La Plata B1900AJK, Argentina
4
Departamento de Química Ambiental, Facultad de Ciencias, Universidad Católica de la Santísima Concepción, Concepción 4090541, Chile
5
Centro de Investigación de Polímeros Avanzados (CIPA), Av. Collao 1202, Concepción 4051381, Chile
6
Laboratorio de Investigaciones Medioambientales de Zonas Áridas (LIMZA), Depto. Ingeniería Mecánica, Facultad de Ingeniería, Universidad de Tarapacá, Arica 1000007, Chile
7
Institute of Condensed Matter and Nanosciences (IMCN), Université Catholique de Louvain, 1348 Louvain-la-Neuve, Belgium
*
Author to whom correspondence should be addressed.
Submission received: 22 January 2022 / Revised: 20 April 2022 / Accepted: 22 April 2022 / Published: 30 April 2022
(This article belongs to the Topic Catalysis for Sustainable Chemistry and Energy)

Abstract

:
A series of tetrabutyl ammonium (TBA) salts of V-included Keggin-type polyoxoanions with W (TBA4PW11V1O40 and TBA5PW10V2O40) and Mo (TBA4PMo11V1O40 and TBA5PMo10V2O40) as addenda atoms were prepared using a hydrothermal method. These synthesized materials were characterized by X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FT-IR), UV-Vis diffuse reflectance (DRS UV-Vis), thermogravimetric analysis (TGA), CHN elemental analysis (EA), inductively coupled plasma spectrometry (ICP-MS), and N2 physisorption techniques to assess their physicochemical/textural properties and correlate them with their catalytic performances. According to FT-IR and DRS UV-Vis, (PVXW(Mo)12−XO40)(3+X)− anions are the main species present in the TBA salts. Additionally, CHN-EA and ICP-MS revealed that the desired stoichiometry was obtained. Their catalytic activities in the liquid-phase aerobic oxidation of benzyl alcohol to benzaldehyde were studied at 5 bar of O2 at 170 °C. Independently of the addenda atom nature, the catalytic activity increased with the number of V in the Keggin anion structure. For both series of catalysts, TBA salts of polyoxometalates with the highest V-substitution degree (TBA5PMo10V2O40 and TBA5PW10V2O40) showed higher activity. The maximum benzyl alcohol conversions obtained were 93% and 97% using (TBA)5PMo10V2O40 and (TBA)5PW10V2O40 as catalysts, respectively. In all the cases, the selectivity toward benzaldehyde was higher than 99%.

1. Introduction

Selective alcohol oxidation is among the most important of oxidation reactions. These reactions are widely used in the chemical industry and in the academic world because carbonyl compounds are important intermediates in chemical synthesis [1]. Conventional processes for alcohol oxidation use some environmentally toxic solvents (e.g., chlorinated) and generally involve harmful inorganic oxidants, such as permanganates, chromates, and peroxides in stoichiometric quantities [2]. Although conventional processes are a viable alternative for the oxidation of alcohols, it is a great challenge to develop selective processes to obtain the desired product. For the reasons stated above, it is imperative to develop new catalytic systems that are eco-friendly and cost-effective for selective alcohol oxidation.
Over the past few years, many catalytic systems have been developed and used for the catalytic oxidation of benzyl alcohol to benzaldehyde, which is a valuable building block for the pharmaceutical, alimentary, and cosmetic industries [3]. To conduct this reaction, vapor-phase and liquid-phase reactions have been studied. Vapor-phase benzyl alcohol catalytic oxidation requires high temperatures, which results in high energy consumption and leads to low selectivity toward benzaldehyde, together with catalyst deactivation [4]. Therefore, liquid-phase oxidation of benzyl alcohol to benzaldehyde is preferable because it can overcome these drawbacks. The choice of oxidant determines the practicability and efficiency of the catalytic systems [5]. In liquid-phase alcohol oxidation, the aerobic system (O2) is widely used as an oxidant because it has been shown to be an ideal oxidant for catalytic oxidation of these substrates [6,7]. Additionally, molecular oxygen O2 has been regarded as the greenest oxidant because it is eco-friendly, inexpensive, easy to handle, and atom efficient, and water is the only byproduct [5].
In recent years, many catalysts have been studied, among which polyoxometalates (POMs) have attracted attention and been revealed to be very active for alcohol oxidation [4]. POMs are inorganic nano-sized anionic transition metal oxide anions with different types of structural architectures [8]. These aggregates consist of d0 metal addenda atoms (usually Mo, W, Nb, etc.) and oxygen arranged in octahedral units (MO6) [9]. POMs with Keggin structures have been intensively studied as green oxidation catalysts for both homogeneous and heterogeneous systems because this catalyst shows both acid and redox properties [10]. The most important characteristic is that their thermal stability and their acidic and redox properties can be tuned by modifying their molecular compositions [11].
Several studies have demonstrated that the catalytic redox activity in Keggin-type polyoxometalates can be enhanced via the substitution of V in place of the Mo or W addenda atom in their primary structure [12,13]. This effect can be explained by considering that M+6 substitution via V+5 increases the POM oxidation potential due to the higher reducibility of vanadium [14]. The role of V substitution in the POM primary structure has been reported in toluene to phenol oxidation, [15], toluene to benzyl alcohol and aldehyde [3], and benzyl alcohol to aldehyde [4]. Frenzel et al. [12] studied the oxidation of sulfides to sulfones using Keggin-type silicopolyoxotungsto compounds (K5(SiVXW11O40) with a variable (1,2) vanadium content and reported that vanadium incorporation improves the catalytic activity under mild reaction conditions.
The main role of countercations in POM chemistry is controlling solubility via the electrostatic stabilization of polyoxoanions [16]. Modification of POMs with organic units has been applied as an effective strategy to improve their catalytic activity, recovery, and reuse. Many organic countercations have been studied for the modulation of POM solubility, such as tetramethyl (TMA), tetrapropyl (TPA), and tetrabutyl (TBA) ammonium salts. Tetrabutyl ammonium cations have been used for many years for POM precipitation into organic solvents [16], showing excellent performance in terms of catalytic oxidation reactions [17].
The points discussed thus far provide background information on the role of vanadium content in the oxidation properties and the catalytic activity of tetraalkylammonium salts of phosphomolybdates and phosphotungstates in catalytic systems.
The aim of this work is the synthesis and characterization of two series of Keggin-type phosphomolybdo (PMoV) and phosphotungstovanadates (PWV) with two different Mo and W replacement degrees (X = 1 and 2) of tetrabutylammonium (TBA) salts, to be used as bulk catalysts in the heterogeneous liquid-phase aerobic oxidation of benzyl alcohol to benzaldehyde.

2. Results and Discussion

2.1. Materials Characterization

The color of (TBA)3PW, (TBA)4PWV and (TBA)5PWV2 samples are shown in Figure 1. In the W-series, the V-unsubstituted (TBA)3PW (a) shows a white color, whilst the V-substituted (TBA)4PWV (b) and (TBA)5PWV2 (c) colorations are yellow and light orange, respectively. Regarding the Mo-series, a similar color change was observed when increasing the amount of V included. This color change is due to the increasing vanadium inclusion in the primary structure of the Keggin-type heteropolyanions. The orange color comes from the intense coloration that V has with an oxidation state of V+5. The colors obtained for these compounds are in accordance with the reported literature [18,19,20,21].

2.2. Characterization

The results of elemental analysis of the materials synthesized in the present study and determined by ICP-MS are shown in Table A1. The results obtained show that such anions have been successfully synthesized following the procedure described above.
The ICP-MS values are in good agreement with the calculated theoretical ones; this fact confirms that the desired Mo(W)/V ratios were obtained. The results of the CHN elemental analysis of the synthesized materials are summarized in Table A2. The elemental analysis confirms that the expected cation/anion ratios for the (TBA)3PW(Mo), (TBA)4PW(Mo)V, and (TBA)5PW(Mo)V2 materials (3, 4, and 5 respectively) were successfully obtained.
Infrared spectra of the (TBA)3PMo, (TBA)4PMoV, (TBA)5PMoV2, (TBA)3PW, (TBA)4PWV, and (TBA)5PWV2 organic salts are shown in Figure 2. These spectra show their main absorption bands in the 800–1100 cm−1 region (which is characteristic of the Keggin structure) and are in agreement with those previously reported in the literature [13,22] for substituted structures. Moreover, the FT-IR spectra exhibit the characteristic vibration bands of organic cation alkyl chains at 1381 and 1470 cm−1 (assigned to the vibration of the C-H bonds) and the band at 1633 cm−1 (assigned to the vibration of H3O+ ions).
Additionally, the not-shown vibration bands at 3440, 2960, and 2875 cm−1 are attributed to H3O+ (the former) and C-H (the latter two).
On the other hand, the FT-IR spectra also show that the maximum of main Keggin band structure shifts to lower frequencies with an increasing V substitution degree in the heteropolyanions (Table A3). These results confirm vanadium atom inclusion in the primary Keggin structure and agree with previous reports [23,24].
The Keggin-type heteropolyanions and their salts display a characteristic XRD diffraction peaks group at the 2θ values between 5° and 10° [13,25]. These characteristic peaks related to the H3PMo12O40 (H3PMo) Keggin structure appears at 6.5 (111), 8.2 (200), and 10.7 (220) of 2θ (hkl) (JCPDS File 01-070-0059). Similar values of 6.9 (010), 8.8 (200), and 9.0 (002) of 2θ (hkl) were reported for H3PW12O40 (H3PW) heteropolyacid (JCPDS File 00-050-0655).
For the TBA-synthesized solids—(TBA)3PMo, (TBA)4PMoV, and (TBA)5PMoV2 and their analogous W-based materials (TBA)3PW, (TBA)4PWV, and (TBA)5PWV2—the XRD patterns show both groups of characteristic diffraction peaks as previously described (Figure 3).
In the case of (TBA)3PMo, the characteristic diffraction peaks related to the Keggin structure are located at 7.1, 8.7, and 10.6 of 2θ. These 2θ values are slightly different from those ascribed to the parent heteropolyacid due to the replacement of H+ by TBA+. However, in the case of (TBA)4PMoV and (TBA)5PMoV2, the increment in the TBA+/anion ratio leads to major differences in peak position and intensity [26,27]. An analogous behavior was observed for the W-series.
The DRS spectra of the Mo-series (TBA)3PMo, (TBA)4PMoV, and (TBA)5PMoV2 organic salts are shown in Figure 4. It has been previously reported that the bands assigned to the charge transfer from bridging or terminal O 2p to W 5d (M-O-M and M-Od, respectively) in the absorption spectrum of unreduced polyanions appear in the range of 200–550 nm [13,18]. The metal M (W, Mo) linked to terminal O has strong double bond character and generates a charge transfer transition in the region of higher energy (210–230 nm). In contrast, the transfer from bridging O 2p to M 5d is observed in the lower energy region (240–550 nm) [28,29].
All of the DRS spectra in Figure 4 display the characteristic bands corresponding to M=O (210–230 nm) and Mo-O-Mo (240–550 nm) electron transfer. The W-series solids (TBA)3PW, (TBA)4PWV, and (TBA)5PWV2 exhibit the same behavior. The absorption edge energy (Eg) in the UV-Vis spectrum reflects the energy required for ligand-to-metal charge transfer (LMCT) [18]. Table 1 shows the absorption edge energy values calculated from the intersection of the extrapolated spectrum with the abscissa axis in the descending part of the curve [30,31].
A continuous diminution of Eg values in the order Eg PM > Eg PMV > Eg PMV2 can be observed. The Eg value decreases in parallel with the increment vanadium atoms in the Keggin anion. The absorption edge energy corresponds to the energy required for the transfer of an electron from the higher energy occupied orbital (HOMO) to the lower energy unoccupied orbit (LUMO) [32]. In Keggin-type POMs, the HOMO is composed of 2p orbitals of the O bridges, while the LUMO is a mixture of d orbitals of the metallic centers of the structure and the 2p orbitals of the neighboring oxygen atoms [33]. These differences in terms of energy have been related to the oxidation potential; therefore, the smaller the energy difference between the HOMO and LUMO is, the greater the wavelength, and the POM will be more easily reduced [29,31]. The (TBA)3PMo and (TBA)3PW structures display higher Eg values for each series, and they decrease in parallel with an increase in the number of vanadium atoms in the Keggin anion. The substitution of an M atom in the Keggin structure does not affect the HOMO energy level because these orbitals are mainly composed of O 2p orbitals. However, the replacement of M by different elements affects the LUMO energy level since they are derived from the d orbitals of M. This indicates that the modification of the LUMO energy level is responsible for the different absorption edge energies of PMV and PMV2 with respect to PM (where M is Mo or W). This result is in line with those previously reported by Barteau et al. [29], showing that V-containing POMs have enhanced redox properties because V stabilizes the LUMO energy level relative to the unsubstituted PM. Furthermore, the redshift of the absorption bands with the increase in V atoms is responsible for the color change observed in the solids (Figure 1).
The TGA analyses up to 240 °C showed no thermal degradation of the V-substituted materials. TGA of the synthesized materials showed a reduction in weight in the range 240–350 °C, corresponding to the organic TBA cation decomposition through a Hoffman’s elimination reaction [18,34] (Figure A1). At temperatures higher than 400 °C, the Keggin anion decomposition begins via the elimination of acidic protons in the form of structural water, as reported in the literature [35,36]. These results allowed us to use the synthesized materials in the aerobic oxidation reaction of benzyl alcohol at temperatures lower than 240 °C.
The N2 adsorption isotherms obtained for the materials are presented in Figure 5, and the main textural property data are summarized in Table 2. The figure shows the main features of type II IUPAC isotherms and the characteristics of non-porous or macro-porous materials [37,38]. The prepared materials display rather low BET area values (in the range of the experimental error ˂ 10 m2g−1), which are in accordance with values reported previously in the literature [21]. It was determined that in both series (Mo and W), the higher the amount of vanadium included, the higher the BET area obtained (Table 2). This effect was reported by Hua et al. [39] and was attributed to an expansion of the latter due to the increase in the alkyl chain length of the countercation, implying that it is more accessible to the substrates. In our case, we could associate it with a higher number of organic cations resulting from a higher number of vanadium atoms [40,41]. The SEM images of the Mo-series (TBA)3PMo, (TBA)4PMoV, and (TBA)5PMoV2 organic salts are shown in Figure 6. In these images, we can observe clear differences in surface morphology between the unsubstituted and V-substituted materials. In the W-series (TBA)3PW, (TBA)4PWV, and (TBA)5PWV2, the same trend was observed. These differences result from the number of organic cations. In addition, it was found that the particles of the unsubstituted materials in each series (TBA)3PMo, (TBA)3PW show a more compact arrangement than the particles of the V-substituted ones. This implies that although the crystals are non-porous, the changes in their aggregation state of the V-substituted materials are responsible for the BET surface area changes.

2.3. Catalytic Benzyl Alcohol Oxidation

In all the liquid-phase aerobic catalytic oxidation of benzyl alcohol reactions, the selectivity was >99%. An example chromatogram is presented in Figure 7, where benzaldehyde is the only reaction product.
The liquid-phase aerobic catalytic oxidation of benzyl alcohol to benzaldehyde conversion results are shown in Figure 8a and Figure A2) (at 4 h of reaction and selectivity at a 100% conversion level) for the Mo and W materials.
The (TBA)3PMo catalyst shows a total conversion of 52% and increases up to 67% for (TBA)4PMoV, achieving a total conversion of 93% for the largest substituted material (TBA)5PMoV2. In the W-series, the (TBA)3PW catalyst exhibits a low conversion (27%). The conversion increases to 82% and 97% in the presence of (TBA)4PWV and (TBA)5PWV2, respectively. It is important to note that the control begins to show conversions after 180 min, reaching a maximum of 18% after 240 min. This may be due to the over-oxidation of the benzyl alcohol; this phenomenon was described by Sankar et al. [42]. Observed rate constants k for benzyl alcohol oxidation are also shown in Figure 8b. Results show that in each series (Mo or W), the observed rate constants become larger as the V-substitution [43] in the solid increases, and all the reactions appear to follow first-order kinetics (Figure A3).
Based on the results presented above, we can determine that the catalytic activity in the oxidation of benzyl alcohol was significantly improved by introducing V into the (TBA)3PMo and (TBA)3PW structure. The conversions of (TBA)4PMV and (TBA)5PMV2 were much higher than those obtained for the V-free phosphomolybdates and phosphotungstates, which implies that V atoms play an important role in the studied reaction [44]. Likewise, the conversions obtained for the inclusion of two V atoms were higher than those obtained for a single V atom. This monotonic increase was previously reported by Li et al. [45] in the oxidation of arenes, and by Yajima et al. [5] in the oxidation of thioamides.
It can be also observed that for some reactions, an appreciable conversion takes place after an induction period. In the Mo-series, (TBA)3PMo shows an induction period of approximately 30 min, (TBA)4PMoV shows 20 min, and (TBA)5PMoV2 (with the largest inclusion of V) shows no induction time. Regarding the W-series, we noticed that for V-unsubstituted (TBA)3PW and mono-substituted (TBA)4PWV, the same induction period of 40 min was observed. No induction time was observed for di-substituted (TBA)5PWV2, as in the Mo-series. Therefore, the induction time tends to decrease with increasing vanadium substitution; this effect should be analyzed. Nomiya et al. [46] studied the induction period in benzene oxidation via V-substituted silicotungstates and found a relationship between this and the formation of vanadium active species in the polyoxotungstate structure. Other authors have reported that this induction period is associated with the formation time of catalytically active species in oxidation reactions of both alcohols (cyclohexanol) [47,48] and olefins [49]. According to Huang et al. [50], in a study of alcohol aerobic oxidation by V-substituted polyoxomolybdates, VO2+ could be the catalytic active species present under reaction conditions [51].
In summary, shortening the induction time and incrementing the maximum conversion upon the addition of V in the Keggin structure were observed for (TBA)3PMo and (TBA)3PW. This effect could be attributed to the phenomenon of the higher the number of vanadium atoms, the higher the concentration and the faster the formation of these species. The net effect of this is a decrease in the induction period and an increase in the maximum conversion rates concerning the unsubstituted forms. Karcz et al. [52], experimentally and via DFT calculations studied the replacement of Co in place of the metal addenda in 12-tungstophosphate and 12-molybdophosphate salts for the aerobic oxidation of cyclooctane, obtaining results that are in accordance as ours.
Moreover, the role of countercations has also been related to the occurrence of an induction time. Guerin et al. [53] analyzed the relationship between the induction times and the use of organic salts as cations in phosphomolybdate salts for the oxidation of cyclooctene, reporting a long induction time (120–240 min) when using TBA salts as cations. This was attributed to the initial mass transport limitations for the oxidant. It is interesting to note that V inclusion leads to an increased TBA as a countercation due to charge compensation. Despite the expected mass transfer limitations, an opposite effect occur; this phenomenon may occur because the magnitude of the effect of V inclusion on the concentration of active species is greater than the effect of the countercation in terms of catalytic activity. Another study [39] analyzed the role of TBA as a countercation in the oxidation of olefins, in which an induction period was observed. It was proposed that the length and structure of the alkyl chains of the countercation play a crucial role in the catalytic performance of POM through the interactions that can be established between these chains and the substrate by means of van der Waals forces (more specifically, in the adsorption and reaction of the substrates) [54]. Another study with TBA salts of (PW10V2O40)5− and (PMo11VO40)4− showed that, despite favoring the occurrence of the induction time, the TBA fulfills the important function of providing structural stability in the oxidation reactions and retaining the original structure of the polyoxoanion for a longer reaction time [46].
The products obtained in the catalytic oxidation of benzyl alcohol depends on the oxidation ability of the catalyst. Thus, if the nature of the catalyst is acidic, the reaction product is diphenylethane (a dimer of benzyl alcohol); if the nature of the material is redox, benzaldehyde is favored [11]. From the latter, we can conclude that liquid-phase aerobic catalytic oxidation of benzyl alcohol with the materials synthesized in this work occurs via a redox pathway. The selectivity values are comparable with those obtained by Abdenatanzi et al. [10] for polyoxotungstate immobilized in ionic liquids, although higher maximum conversions were obtained in our case. Therefore, for the synthesized pure and substituted Mo and W catalysts, the obtained results reveal that they are very selective in benzyl alcohol oxidation toward benzaldehyde, without forming over-oxidized products, the Mo-based catalysts showing slightly shorter induction times than their W-based counterparts. Regarding selectivity, no appreciable differences were observed. Similar results were obtained in the literature [53] for cyclooctene oxidation by butylpiridinium salts of PW12O40 and PMo12O40.

2.4. Catalyst Reuse

Reusability is a valuable characteristic of a solid heterogeneous catalyst. For the catalyst reuse study, we used one of the catalysts that presented the highest conversions [(TBA)5PMoV2] for liquid-phase aerobic catalytic oxidation of benzyl alcohol reaction. To evaluate the reusability of (TBA)5PMoV2, the catalyst was recovered and reused in the same reaction conditions. The results obtained for the three reuses at 4 h, 5 h, 6 h, and 7 h are summarized in Figure 9.
From these results, we can conclude that conversion values obtained when reusing the (TBA)5PMoV2 catalyst for three consecutive times are quite similar to those obtained when using a fresh catalyst. In all the cases, the selectivity values achieved were higher than 99%.
The FT-IR spectrum and the XRD diffractogram of the (TBA)5PMoV2 reused three times (Figure 10a,b, respectively) reveal that slight differences in the XRD diagram and in the FT-IR spectrum were detected, compared to the fresh sample. The slight differences observed between the diffraction patterns of fresh and reused POM may correspond to the loss and/or modification of crystalline phases. However, since these phases are not catalytically active, there are no appreciable differences in the conversion values in catalyst reuse. The SEM micrograph of the reused solid (Figure 10c) shows that the surface morphology of the catalyst does not undergo appreciable changes after the reuse experiments, which reinforces the information obtained by FT-IR and XRD.
The excellent reusability observed in these catalysts agrees with the catalytic mechanism reported in [11]. The slower step is based on the abstraction of the benzyl proton by the adjacent oxygen, and the subsequent rearrangement of the electrons results in the subsequent formation of the carbonyl group with the bridging oxygen. In this step, the desorption of the benzaldehyde molecule generates an oxygen vacancy, leaving the catalyst partially reduced. The last step is the reoxidation of the catalyst with molecular oxygen [55,56]. Thus, it is possible to suggest that V-substituted Mo- and W-type polyoxometalates have similar catalytic behavior, which would explain the high reusability.

2.5. Leaching Study

The benzyl alcohol conversion (1.5%) obtained in the leaching test (see Experimental 3.7) was comparable to the control experiment value at an equivalent time. These experiments proved that no detectable leaching of the catalytically active species into the solvent took place, allowing us to discard homogeneous phase reactions.
Finally, for comparative purposes, Table 3 summarizes some recently reported catalytic systems employed in the catalytic oxidation of benzyl alcohol. From these figures, we can assume that our catalytic system is competitive.

3. Materials and Methods

3.1. Synthesis of Heteropolycompounds [N(Butyl)4]3[PMo12O40]-TBA3PMo and [N(Butyl)4]3[PW12O40]-TBA3PW

Commercial heteropolyacids H3PMo12O40 (99.9%, Merck, Darmstadt, Germany) and H3PW12O40 (99.9%, Merck, Darmstadt, Germany) were precipitated as their corresponding organic salts of tetrabutylammonium (TBABr, 99.5%, Merck, Darmstadt, Germany) according to a previously reported procedure [63].

3.2. Synthesis of [N(Butyl)4](3+x)[PVxMo12−xO40] with x = 1 (TBA)4PMoV and x = 2 (TBA)5PMoV2

The TBA salts of the Keggin-type phosphomolybdovanadates [PMo11VO40]4− (PMoV) and [PMo10V2O40]5− (PMoV2) were prepared using a hydrothermal process [13,64]. In a typical synthesis for the V-mono-substituted type, a stoichiometric mixture of 0.01 mol H3PO4 (85%, Merck, Darmstadt, Germany), 0.005 mol V2O5 (99%, Merck, Darmstadt, Germany), and 0.11 mol MoO3 (99,5%, Merck, Darmstadt, Germany) was suspended in 150 mL of distilled water. The mixture was then heated and stirred at 80 °C and 800 rpm on a hotplate for 4 h. After cooling to room temperature and removing the insoluble molybdates and vanadates via vacuum filtration, the heteropolyacid solution was again heated to 60 °C, and tetrabutylammonium bromide (0.012 mol) was added. Precipitation of the organic salt was observed immediately (at 60 °C). The obtained organic salt was filtered at room temperature and dried under a vacuum to a constant weight (±0.002 g). For the synthesis of the V-di-substituted type, 0.01 mol of V2O5 was used.

3.3. Synthesis of [N(Butyl)4](3+x)[PVxW12−xO40] with x = 1 (TBA)4PWV and x = 2 (TBA)5PWV2)

Phosphotungstovanadates [PW11VO40]4− (PWV) and [PW10V2O40]5− (PWV2) were also prepared using a hydrothermal process [18,25]. The first step was to prepare a stock solution of vanadium (V) by mixing 0.25 mol NH4VO3 (99%, Merck, Germany) with 0.50 mol NaOH (99.5%, Merck, Germany) in 500 mL of distilled water. Then, 0.0025 mol NaH2PO4·12H2O (99.9%, Merck, Germany) was added to an aqueous solution of 0.019 mol Na2WO4·2H2O (99%, Merck, Germany) with continuous stirring, and the pH was adjusted to 2.8 with HCl (36% v/v, Merck, Germany) dropwise, with 3.75 (PWV) or 15.0 mL (PWV2) of the vanadium stock solution added. This mixture was then heated at 80 °C for 4 h on a hot plate. The final obtained solution was cooled and filtered to remove any insoluble salts. Finally, it was heated to 60 °C and tetrabutylammonium bromide (0.012 mol) was added under continuous stirring; precipitation of the organic salt was observed immediately (at 60 °C). The organic salts obtained were finally filtered at room temperature and dried under a vacuum to a constant weight.

3.4. Characterization

The XRD patterns were recorded at room temperature in a D8 ADVANCE (Bruker, Karlsruhe, Germany) diffractometer with Bragg–Brentano geometry using a LYNXEYE XE-T detector and copper Kα radiation (λ = 1.5406 Å) for a 2θ range of 5–80° (step = 0.015°, time per step = 0.15 s). The X-ray source was operated with a tension of 40 kV and a current of 30 mA. The FT-IR spectra were recorded using a Nicolet Nexus 470 spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) in the 400–4000 cm−1 wavenumber range at room temperature, and KBr tablets were used. The DRS UV-Vis spectra were recorded on a UV-3600 Plus UV-Vis-NIR Spectrophotometer (Shimadzu, Kyoto, Japan) with a Harrick single-beam Praying Mantis Diffuse Reflectance collection system. The spectra were then recorded at room temperature in the 190–1000 nm range. A Spectralon® Diffuse Reflectance Standard was used to measure the background spectra. The DRS UV-Vis spectra were background corrected and normalized by the most intense absorption band, and the Kubelka–Munk function was used to display the data. A TGA analysis was performed on a TG 209F1 Iris (NETZSCH, Selb, Germany) thermogravimetric analyzer under an N2 atmosphere in the temperature range of 25–550 °C using 10–25 mg of the solid and a heating rate of 10 °C min−1. The CHN elemental analysis was performed on an EA 2400 series II (Perkin Elmer, Waltham, MA, USA); measurements were carried out with samples of 10–25 mg in weight. Inductively coupled plasma mass spectrometry (ICP-MS) measurements for Mo, W, and V determination in the materials were performed using an iCAP Q series ICP-MS (Thermo Scientific, Waltham, MA, USA) spectrometer. Specific areas, the pore volume, and the pore diameter were evaluated via the Brunauer–Emmet–Teller (BET) method using the nitrogen adsorption isotherms obtained on a ASAP 2010 (Micromeritics Instrument Corporation, Norcross, GA, USA) apparatus at −196 °C in the relative pressure range of 0.01–0.99 (P/P0). The samples were previously pre-treated at 150 °C for 3 h. The specific surface area (SSA) was calculated from the adsorption isotherm branch using the BET method.

3.5. Catalytic Benzyl Alcohol Oxidation

The catalytic activity of the obtained solids was examined using benzyl alcohol as the reaction substrate. The tests were performed in an O2 atmosphere in a semi-batch Parr-type reactor with a methanol–water solvent (90:10 v/v) at 170 °C for 4 h. In a typical reaction, 1 g (0.01 mol) of benzyl alcohol, 100 mg of the catalyst, and 20 mL of the solvent were added to the reactor. Subsequently, in order to avoid contamination with other gases, after the reactor was sealed, it was purged three times with O2 (99.5%, Linde, Santiago, Chile) and pressurized with 5 bar of the same gas. Finally, the pressurized reactor was heated to 170 °C (10 °C min−1 heating ramp), and the O2 pressure was adjusted to 5 bar. The selected agitation speed was 600 rpm to avoid undesirable diffusion control during the catalytic performance evaluation. The study was then conducted for all the catalysts by taking samples at different reaction times for 4 h. All the reaction products were in line with the commercial standards. The reaction products were identified and quantified by using a FLEXARTM HLPC-DAD (Perkin Elmer Inc., Waltham, MA, USA) with a mobile phase composition of 0.1% glacial acetic acid (A) and acetonitrile (B) with a flow rate of 1.0 mL min−1. The column used was an Inertsil C18 ODS-3 (250 mm × 4.6 mm, 5 μm) (GL Sciences Inc., Torrance, CA, USA), and the HPLC-DAD detection wavelength was 254 nm. Finally, the reaction products were confirmed by GC-MS using a model 7890A GC system (Agilent Technologies Inc., Wilmington, DE, USA) interfaced to a single quadrupole mass-selective detector 5975C (triple-axis detector).
The conversion for benzyl alcohol (XBzOH) and the selectivity toward the products (Sproduct) were calculated by using calibration curves as follows:
X BzOH ( % ) = [ BzOH ] i [ BzOH ] t [ BzOH ] i × 100
S product ( % ) =   [ product ] t [ BzOH ] i [ BzOH ] t × 100  
The control experiment was performed with the same as the typical reaction mentioned above, but without any catalyst addition.

3.6. Catalyst Reuse

The reusability of the catalysts was evaluated in three consecutive runs, using the same reaction conditions used in benzyl alcohol oxidation. The catalyst was isolated from the reaction mixture after each run, washed twice with ethanol (2 mL), dried under vacuum at 25 °C until constant weight, and then reused in a new experiment. For reusability experiments, the conversion and selectivity were evaluated at 4, 5, 6, and 7 h with the purpose of evaluating the reproducibility of the experiment. The solid obtained at the end of the experiments was characterized by XRD, FT-IR and SEM.

3.7. Leaching Studies

To assess the possibility of catalyst leaching to the solvent, ruling it out it or not, an additional experiment was carried out. In this experiment, 100 mg of the catalyst and 20 mL of the solvent were added to the reactor and heated to 170 °C. The O2 pressure was adjusted to 5 bar and the reaction time was 3 h. Subsequently, the solid was separated by filtration and the solvent was used in a new reaction with 0.01 mol of benzyl alcohol without catalyst addition, under the same conditions used in the benzyl alcohol oxidation test (Section 3.5).

4. Conclusions

In this work, a series of tetrabutyl ammonium (TBA) salts of V-included Keggin-type polyoxoanions with W (TBA4PW11V1O40 and TBA5PW10V2O40) and Mo (TBA4PMo11V1O40 and TBA5PMo10V2O40) as addenda atoms were prepared and characterized by XRD, FT-IR, and DRS UV-Vis techniques.
The catalytic performance in the selective liquid-phase aerobic oxidation of benzyl alcohol, using O2 as a green oxidant agent, indicates that in both the Mo and W series, the replacement of addenda atoms by V enhances the catalytic activity with respect to non-substituted atoms. Under the reaction conditions used, the di-substituted (TBA)5PWV2 and (TBA)5PMoV2 exhibited higher catalytic activity for the benzyl alcohol oxidation, with total conversions of 97% and 93%, respectively, and a selectivity towards benzaldehyde > 99%. These catalysts also proved to be reusable under these reaction conditions. The findings of this study on the effect of V-inclusion polyoxotugsto- and polyoxomolybdovanadates with Keggin structures in the catalytic oxidation of benzyl alcohol will lead to a better understanding of the mechanisms involved in this reaction.

Author Contributions

J.D.: Conduction of the investigation process, performance of the experiments, and preparation of the published work. L.R.P.: preparation of the published work. G.P.: provision of instrumentation, review and edition of the published work. C.H.C.: preparation of the published work and acquisition of financial support. L.A.: application of graphic tools and redaction. R.B.: acquisition of financial support. R.R.: provision of instrumentation. A.H.: Methodology. E.M.G.: provision of instrumentation. D.C.: acquisition of financial support; leadership responsibility for the research activity, planning, and execution; and preparation of the published work. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank Millennium Science Initiative Program—NCN2021_090, Project National Agency for Research and Development ANID-FONDECYT 1201895, Chilean ANID Doctorate Grant/2018–21181942, Doctoral Thesis in Productive Area N° T7819120007, and ANID/Programa de Cooperación Internacional/REDES180165 FONDAP SERC-CHILE 1510019 for financially supporting this research.

Data Availability Statement

Not applicable.

Conflicts of Interest

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

Appendix A

Table A1. Metal analysis of pure and substituted Mo and W catalysts determined by ICP-MS.
Table A1. Metal analysis of pure and substituted Mo and W catalysts determined by ICP-MS.
CatalystM a
(wt%) Exp. b
V
(wt%) Exp.
M a
(wt%) Calc. c
V
(wt%) Calc.
M a mmol
Exp.
V
mmol
Exp.
M a/V
Ratio
Exp.
M a/V
Ratio
Calc.
(TBA)3PMo43.7-45.1-4.6---
(TBA)4PMoV39.11.938.41.84.10.410.911.0
(TBA)5PMoV235.13.932.63.53.70.84.85.0
(TBA)3PW60.8-61.2-3.4---
(TBA)4PWV56.71.554.51.43.10.310.511.0
(TBA)5PWV250.12.848.12.72.70.55.05.0
a M = Mo or W; b experimental values; c calculated values as anhydrous salts.
Table A2. CHN analysis of the pure and substituted Mo and W catalysts by elemental analysis.
Table A2. CHN analysis of the pure and substituted Mo and W catalysts by elemental analysis.
CatalystC (wt%)H (wt%)N
(wt%)
Cation
(mmol × 103) a
Anion
(mmol × 103) b
Cation/Anion
Exp. c
Cation/Anion
Theory d
(TBA)3PMo23.14.41.63.01.03.13
(TBA)4PMoV25.24.71.82.00.53.94
(TBA)5PMoV229.15.32.02.00.44.95
(TBA)3PW16.12.91.02.10.73.03
(TBA)4PWV17.74.51.32.50.64.44
(TBA)5PWV220.63.91.51.60.34.85
a cation corresponds to a TBA salt; b anion corresponds to a (PMo(12−x)Vx O40)(3+x)− or (PW(12−x)VxO40)(3+x)− moiety; c experimental value; d theoretical values calculated as anhydrous salts.
Regarding (TBA)3PMo, the FT-IR spectrum shows bands at 1064, 960, 880, and 809 cm−1, assigned to stretching vibrations of P-Oa, Mo-Od, Mo-Ob-Mo, and Mo-Oc-Mo, respectively, where Oa surrounds the central P heteroatom and connects P with Mo, Ob connects two Mo3O13 triplets (three edge-shared octahedra) via corner sharing, Oc connects two Mo3O13 triplets via edge sharing, and the terminal Od is bonded to a single Mo atom. Likewise, (TBA)3PW shows bands at 1081, 981, 892, and 816 cm−1, which are also assigned to the P-Oa, W-Od, W-Ob-W, and W-Oc-W stretching bands, respectively. The frequency values for all the materials are summarized in Table A3.
Table A3. Characteristic FT-IR bands of the pure and substituted Mo and W catalysts.
Table A3. Characteristic FT-IR bands of the pure and substituted Mo and W catalysts.
Catalystν P-Oa
(cm−1)
ν Mo-Od
(cm−1)
ν Mo-Ob-Mo
(cm−1)
ν Mo-Oc-Mo
(cm−1)
(TBA)3PMo1064960880809
(TBA)4PMoV1061959879806
(TBA)5PMoV21060954877805
(TBA)3PW1081981892816
(TBA)4PWV1068968889813
(TBA)5PWV21067967889812
Figure A1. TGA diagrams of pure and V-substituted Mo (a) and W (b) Keggin-type polyoxometalates.
Figure A1. TGA diagrams of pure and V-substituted Mo (a) and W (b) Keggin-type polyoxometalates.
Catalysts 12 00507 g0a1
Figure A2. Total conversion (%) of the liquid-phase aerobic oxidation of benzyl alcohol over time for the pure and substituted Mo (a) and W (b) catalysts.
Figure A2. Total conversion (%) of the liquid-phase aerobic oxidation of benzyl alcohol over time for the pure and substituted Mo (a) and W (b) catalysts.
Catalysts 12 00507 g0a2
Figure A3. First-order kinetic analysis plot for the liquid-phase aerobic oxidation reaction of benzyl alcohol over time for the pure and substituted Mo and W catalysts.
Figure A3. First-order kinetic analysis plot for the liquid-phase aerobic oxidation reaction of benzyl alcohol over time for the pure and substituted Mo and W catalysts.
Catalysts 12 00507 g0a3

References

  1. Ilyas, M.; Siddique, M.; Saeed, M. Liquid-phase aerobic oxidation of benzyl alcohol catalyzed by mechanochemically synthesized manganese oxide. Chin. Sci. Bull. 2013, 58, 2354–2359. [Google Scholar] [CrossRef] [Green Version]
  2. Adnan, R.H.; Andersson, G.G.; Polson, M.I.J.; Metha, G.F.; Golovko, V.B. Factors influencing the catalytic oxidation of benzyl alcohol using supported phosphine-capped gold nanoparticles. Catal. Sci. Technol. 2015, 5, 1323–1333. [Google Scholar] [CrossRef] [Green Version]
  3. Djaouida, A.; Sadia, M.; Smaïn, H. Direct benzyl alcohol and benzaldehyde synthesis from toluene over Keggin-type polyoxometalates catalysts: Kinetic and mechanistic studies. J. Chem. 2019, 2019, 9521529. [Google Scholar] [CrossRef] [Green Version]
  4. Jing, L.; Shi, J.; Zhang, F.; Zhong, Y.; Zhu, W. Polyoxometalate-based amphiphilic catalysts for selective oxidation of benzyl alcohol with hydrogen peroxide under organic solvent-free conditions. Ind. Eng. Chem. Res. 2013, 52, 10095–10104. [Google Scholar] [CrossRef]
  5. Yajima, K.; Yamaguchi, K.; Mizuno, N. Facile access to 3,5-symmetrically disubstituted 1,2,4-thiadiazoles through phosphovanadomolybdic acid catalyzed aerobic oxidative dimerization of primary thioamides. Chem. Commun. 2014, 50, 6748–6750. [Google Scholar] [CrossRef]
  6. Wang, S.; Li, S.; Shi, R.; Zou, X.; Zhang, Z.; Fu, G.; Li, L.; Luo, F. A nanohybrid self-assembled from exfoliated layered vanadium oxide nanosheets and Keggin Al13 for selective catalytic oxidation of alcohols. Dalton Trans. 2020, 49, 2559–2569. [Google Scholar] [CrossRef]
  7. Narkhede, N.; Patel, A.; Singh, S. Mono lacunary phosphomolybdate supported on MCM-41: Synthesis, characterization and solvent free aerobic oxidation of alkenes and alcohols. Dalton Trans. 2014, 43, 2512–2520. [Google Scholar] [CrossRef]
  8. Huang, J.B.; Wu, S.b.; Cheng, H.; Lei, M.; Liang, J.J.; Tong, H. Theoretical study of bond dissociation energies for lignin model compounds. Ranliao Huaxue Xuebao/J. Fuel Chem. Technol. 2015, 43, 429–436. [Google Scholar] [CrossRef]
  9. Grigoriev, V.A.; Hill, C.L.; Weinstock, I.A. Polyoxometalate Oxidation of Phenolic Lignin Models. ACS Symp. Ser. 2009, 785, 297–312. [Google Scholar] [CrossRef]
  10. Abednatanzi, S.; Abbasi, A.; Masteri-Farahani, M. Immobilization of catalytically active polyoxotungstate into ionic liquid-modified MIL-100(Fe): A recyclable catalyst for selective oxidation of benzyl alcohol. Catal. Commun. 2017, 96, 6–10. [Google Scholar] [CrossRef]
  11. Rao, P.S.N.; Parameswaram, G.; Rao, A.V.P.; Lingaiah, N. Influence of cesium and vanadium contents on the oxidation functionalities of heteropoly molybdate catalysts. J. Mol. Catal. A Chem. 2015, 399, 62–70. [Google Scholar] [CrossRef]
  12. Frenzel, R.A.; Palermo, V.; Sathicq, A.G.; Elsharif, A.M.; Luque, R.; Pizzio, L.R.; Romanelli, G.P. A green and reusable catalytic system based on silicopolyoxotungstovanadates incorporated in a polymeric material for the selective oxidation of sulfides to sulfones. Microporous Mesoporous Mater. 2021, 310, 110584. [Google Scholar] [CrossRef]
  13. Villabrille, P.; Romanelli, G.; Gassa, L.; Vázquez, P.; Cáceres, C. Synthesis and characterization of Fe- and Cu-doped molybdovanadophosphoric acids and their application in catalytic oxidation. Appl. Catal. A Gen. 2007, 324, 69–76. [Google Scholar] [CrossRef]
  14. Lee, J.K.; Melsheimer, J.; Berndt, S.; Mestl, G.; Schlögl, R.; Köhler, K. Transient responses of the local electronic and geometric structures of vanado-molybdo-phoshate catalysts H3+nPVnMo12−nO40 in selective oxidation. Appl. Catal. A Gen. 2001, 214, 125–148. [Google Scholar] [CrossRef]
  15. Yuan, C.; Gao, X.; Pan, Z.; Li, X.; Tan, Z. Molybdovanadophosphoric anion ionic liquid as a reusable catalyst for solvent-free benzene oxidation to phenol by H2O2. Catal. Commun. 2015, 58, 215–218. [Google Scholar] [CrossRef]
  16. Misra, A.; Kozma, K.; Streb, C.; Nyman, M. Beyond Charge Balance: Counter-Cations in Polyoxometalate Chemistry. Angew. Chem. Int. Ed. 2020, 59, 596–612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Dornan, L.M.; Muldoon, M.J. A highly efficient palladium(II)/polyoxometalate catalyst system for aerobic oxidation of alcohols. Catal. Sci. Technol. 2015, 5, 1428–1432. [Google Scholar] [CrossRef]
  18. Frenzel, R.A.; Romanelli, G.P.; Pizzio, L.R. Novel catalyst based on mono- and di-vanadium substituted Keggin polyoxometalate incorporated in poly (acrylic acid-co-acrylamide) polymer for the oxidation of sulfides. Mol. Catal. 2018, 457, 8–16. [Google Scholar] [CrossRef]
  19. Pinto, T.V.; Fernandes, D.M.; Guedes, A.; Cardoso, N.; Durães, N.F.; Silva, C.; Pereira, C.; Freire, C. Photochromic polypropylene fibers based on UV-responsive silica@phosphomolybdate nanoparticles through melt spinning technology. Chem. Eng. J. 2018, 350, 856–866. [Google Scholar] [CrossRef]
  20. Babu, B.H.; Parameswaram, G.; Kumar, A.S.H.; Prasad, P.S.; Lingaiah, N. Vanadium containing heteropoly molybdates as precursors for the preparation of MoVP oxides supported on alumina catalysts for ammoxidation of m-xylene. Appl. Catal. A Gen. 2012, 445–446, 339–345. [Google Scholar] [CrossRef]
  21. Palermo, V.; Sathicq, G.; Constantieux, T.; Rodríguez, J.; Vázquez, P.G.; Romanelli, G.P. First Report about the Use of Micellar Keggin Heteropolyacids as Catalysts in the Green Multicomponent Synthesis of Nifedipine Derivatives. Catal. Lett. 2016, 146, 1634–1647. [Google Scholar] [CrossRef]
  22. Gong, Y.; Hu, C.; Li, H.; Tang, W.; Huang, K.; Hou, W. Synthesis, crystal structure and calcination of three novel complexes based on 2-aminopyridine and polyoxometalates. J. Mol. Struct. 2005, 784, 228–238. [Google Scholar] [CrossRef]
  23. Rocchiccioli-Deltcheff, C.; Fournier, M. Catalysis by polyoxometalates. Part 3.—Influence of vanadium(V) on the thermal stability of 12-metallophosphoric acids from in situ infrared studies. J. Chem. Soc. Faraday Trans. 1991, 87, 3913–3920. [Google Scholar] [CrossRef]
  24. Fournier, M.; Thouvenot, R.; Rocchiccioli-Deltcheff, C. Catalysis by polyoxometalates part 1.—Supported polyoxoanions of the Keggin structure: Spectroscopic study (IR, Raman, UV) of solutions used for impregnation. J. Chem. Soc. Faraday Trans. 1991, 87, 349–356. [Google Scholar] [CrossRef]
  25. Ueda, T.; Komatsu, M.; Hojo, M. Spectroscopic and voltammetric studies on the formation of Keggin-type V(V)-substituted tungstoarsenate(V) and-phosphate(V) complexes in aqueous and aqueous-organic solutions. Inorg. Chim. Acta 2003, 344, 77–84. [Google Scholar] [CrossRef]
  26. Tang, S.; Wu, W.; Fu, Z.; Zou, S.; Liu, Y.; Zhao, H.; Kirk, S.R.; Yin, D. Vanadium-Substituted Tungstophosphoric Acids as Efficient Catalysts for Visible-Light-Driven Oxygenation of Cyclohexane by Dioxygen. ChemCatChem 2015, 7, 2637–2645. [Google Scholar] [CrossRef]
  27. Ressler, T.; Timpe, O.; Girgsdies, F.; Wienold, J.; Neisius, T. In situ investigations of the bulk structural evolution of vanadium-containing heteropolyoxomolybdate catalysts during thermal activation. J. Catal. 2005, 231, 279–291. [Google Scholar] [CrossRef] [Green Version]
  28. Viswanadham, B.; Srikanth, A.; Chary, K.V.R. Characterization and reactivity of 11-molybdo-1-vanadophosphoric acid catalyst supported on zirconia for dehydration of glycerol to acrolein. J. Chem. Sci. 2014, 126, 445–454. [Google Scholar] [CrossRef]
  29. Barteau, K.P.; Lyons, J.E.; Song, I.K.; Barteau, M.A. UV–visible spectroscopy as a probe of heteropolyacid redox properties: Application to liquid phase oxidations. Top. Catal. 2006, 41, 55–62. [Google Scholar] [CrossRef]
  30. Abdelghany, A.; Abdelrazek, E.; Badr, S.; Morsi, M. Effect of gamma-irradiation on (PEO/PVP)/Au nanocomposite: Materials for electrochemical and optical applications. Mater. Des. 2016, 97, 532–543. [Google Scholar] [CrossRef]
  31. Tan, C.; Bu, W. Synthesis and energy band characterization of hybrid molecular materials based on organic–polyoxometalate charge-transfer salts. J. Solid State Chem. 2014, 219, 93–98. [Google Scholar] [CrossRef]
  32. Yamase, T. Photo- and Electrochromism of Polyoxometalates and Related Materials. Chem. Rev. 1998, 98, 307–326. [Google Scholar] [CrossRef] [PubMed]
  33. Eguchi, K.; Seiyama, T.; Yamazoe, N.; Katsuki, S.; Taketa, H. Electronic Structures of XMo12040 Heteropolyanions (X = P, As, Si, and Ge) and Their Reduction Behavior. J. Catal. 1988, 111, 336–344. [Google Scholar] [CrossRef]
  34. Gamelas, J.; Couto, F.A.; Trovão, M.N.; Cavaleiro, A.; Cavaleiro, J.; de Jesus, J.D. Investigation of the thermal decomposition of some metal-substituted Keggin tungstophosphates. Thermochim. Acta 1999, 326, 165–173. [Google Scholar] [CrossRef]
  35. Song, Y.-F.; Long, D.-L.; Cronin, L. Hybrid polyoxometalate clusters with appended aromatic platforms. CrystEngComm 2009, 12, 109–115. [Google Scholar] [CrossRef]
  36. Jing, F.; Katryniok, B.; Dumeignil, F.; Bordes-Richard, E.; Paul, S. Catalytic selective oxidation of isobutane to methacrylic acid on supported (NH4)3HPMo11VO40 catalysts. J. Catal. 2014, 309, 121–135. [Google Scholar] [CrossRef]
  37. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  38. Sharma, M.; Saikia, G.; Ahmed, K.; Gogoi, S.R.; Puranik, V.G.; Islam, N.S. Vanadium-based polyoxometalate complex as a new and efficient catalyst for phenol hydroxylation under mild conditions. New J. Chem. 2018, 42, 5142–5152. [Google Scholar] [CrossRef]
  39. Hua, L.; Qiao, Y.; Yu, Y.; Zhu, W.; Cao, T.; Shi, Y.; Li, H.; Feng, B.; Hou, Z. A Ti-substituted polyoxometalate as a heterogeneous catalyst for olefin epoxidation with aqueous hydrogen peroxide. New J. Chem. 2011, 35, 1836–1841. [Google Scholar] [CrossRef]
  40. Arichi, J.; Eternot, M.; Louis, B. Synthesis of V-containing Keggin polyoxometalates: Versatile catalysts for the synthesis of fine chemicals? Catal. Today 2008, 138, 117–122. [Google Scholar] [CrossRef]
  41. Moffat, J.; McMonagle, J.; Taylor, D. Microporous heteropoly oxometalate heterogeneous catalysis. Solid State Ion. 1988, 26, 101–108. [Google Scholar] [CrossRef]
  42. Sankar, M.; Nowicka, E.; Carter, E.; Murphy, D.; Knight, D.W.; Bethell, D.; Hutchings, G.J. The benzaldehyde oxidation paradox explained by the interception of peroxy radical by benzyl alcohol. Nat. Commun. 2014, 5, 3332. [Google Scholar] [CrossRef] [PubMed]
  43. Khenkin, A.; Neumann, R. Low-Temperature Activation of Dioxygen and Hydrocarbon Oxidation Catalyzed by a Phosphovanadomolybdate: Evidence for a Mars–van Krevelen Type Mechanism in a Homogeneous Liquid Phase. Angew. Chem. Int. Ed. 2000, 39, 4088–4090. [Google Scholar] [CrossRef]
  44. Wang, S.-S.; Yang, G.-Y. Recent Advances in Polyoxometalate-Catalyzed Reactions. Chem. Rev. 2015, 115, 4893–4962. [Google Scholar] [CrossRef] [PubMed]
  45. Li, C.; Suzuki, K.; Yamaguchi, K.; Mizuno, N. Phosphovanadomolybdic acid catalyzed direct C–H trifluoromethylation of (hetero)arenes using NaSO2CF3 as the CF3 source and O2 as the terminal oxidant. New J. Chem. 2017, 41, 1417–1420. [Google Scholar] [CrossRef]
  46. Nomiya, K.; Nemoto, Y.; Hasegawa, T.; Matsuoka, S. Multicenter active sites of vanadium-substituted polyoxometalate catalysts on benzene hydroxylation with hydrogen peroxide and two reaction types with and without an induction period. J. Mol. Catal. A Chem. 2000, 152, 55–68. [Google Scholar] [CrossRef]
  47. Ni, L.; Patscheider, J.; Baldridge, K.K.; Patzke, G.R. New Perspectives on Polyoxometalate Catalysts: Alcohol Oxidation with Zn/Sb-Polyoxotungstates. Chem. Eur. J. 2012, 18, 13293–13298. [Google Scholar] [CrossRef]
  48. Mizuno, N.; Kamata, K. Catalytic oxidation of hydrocarbons with hydrogen peroxide by vanadium-based polyoxometalates. Coord. Chem. Rev. 2011, 255, 2358–2370. [Google Scholar] [CrossRef]
  49. Kamata, K.; Kotani, M.; Yamaguchi, K.; Hikichi, S.; Mizuno, N. Olefin Epoxidation with Hydrogen Peroxide Catalyzed by Lacunary Polyoxometalate [γ-SiW10O34(H2O)2]4−. Chem. Eur. J. 2006, 13, 639–648. [Google Scholar] [CrossRef]
  50. Huang, X.; Zhang, X.; Zhang, D.; Yang, S.; Feng, X.; Li, J.; Lin, Z.; Cao, J.; Pan, R.; Chi, Y.; et al. Binary Pd-Polyoxometalates and Isolation of a Ternary Pd-V-Polyoxomolybdate Active Species for Selective Aerobic Oxidation of Alcohols. Chem. Eur. J. 2014, 20, 2557–2564. [Google Scholar] [CrossRef]
  51. Ivanchikova, I.D.; Maksimchuk, N.V.; Maksimovskaya, R.I.; Maksimov, G.M.; Kholdeeva, O.A. Highly Selective Oxidation of Alkylphenols to p-Benzoquinones with Aqueous Hydrogen Peroxide Catalyzed by Divanadium-Substituted Polyoxotungstates. ACS Catal. 2014, 4, 2706–2713. [Google Scholar] [CrossRef]
  52. Karcz, R.; Niemiec, P.; Pamin, K.; Połtowicz, J.; Kryściak-Czerwenka, J.; Napruszewska, B.D.; Michalik-Zym, A.; Witko, M.; Tokarz-Sobieraj, R.; Serwicka, E.M. Effect of cobalt location in Keggin-type heteropoly catalysts on aerobic oxidation of cyclooctane: Experimental and theoretical study. Appl. Catal. A Gen. 2017, 542, 317–326. [Google Scholar] [CrossRef]
  53. Guérin, B.; Fernandes, D.M.; Daran, J.-C.; Agustin, D.; Poli, R. Investigation of induction times, activity, selectivity, interface and mass transport in solvent-free epoxidation by H2O2 and TBHP: A study with organic salts of the [PMo12O40]3− anion. New J. Chem. 2013, 37, 3466–3475. [Google Scholar] [CrossRef]
  54. Neumann, R. Activation of Molecular Oxygen, Polyoxometalates, and Liquid-Phase Catalytic Oxidation. Inorg. Chem. 2010, 49, 3594–3601. [Google Scholar] [CrossRef]
  55. Albert, J.; Lüders, D.; Bösmann, A.; Guldi, D.M.; Wasserscheid, P. Spectroscopic and electrochemical characterization of heteropoly acids for their optimized application in selective biomass oxidation to formic acid. Green Chem. 2013, 16, 226–237. [Google Scholar] [CrossRef]
  56. Bertleff, B.; Claußnitzer, J.; Korth, W.; Wasserscheid, P.; Jess, A.; Albert, J. Extraction Coupled Oxidative Desulfurization of Fuels to Sulfate and Water-Soluble Sulfur Compounds Using Polyoxometalate Catalysts and Molecular Oxygen. ACS Sustain. Chem. Eng. 2017, 5, 4110–4118. [Google Scholar] [CrossRef]
  57. Santonastaso, M.; Freakley, S.J.; Miedziak, P.J.; Brett, G.L.; Edwards, J.K.; Hutchings, G.J. Oxidation of Benzyl Alcohol using in Situ Generated Hydrogen Peroxide. Org. Process Res. Dev. 2014, 18, 1455–1460. [Google Scholar] [CrossRef]
  58. Saffari, N.S.; Aghabarari, B.; Javaheri, M.; Khanlarkhani, A.; Martinez-Huerta, M.V. Transforming Waste Clamshell into Highly Selective Nanostructured Catalysts for Solvent Free Liquid Phase Oxidation of Benzyl Alcohol. Catalysts 2022, 12, 155. [Google Scholar] [CrossRef]
  59. Xuan, Y.; Haojie, L.; Tianhao, Y.; Zhexin, X.; Qiqi, W.; Zhengping, Z. Preparation and Application of Metal Oxides/SBA-15 Mesoporous Composites as Catalysts for Selective Oxidation of benzyl alcohol. Int. J. Electrochem. Sci. 2022. [Google Scholar] [CrossRef]
  60. Chen, G.; Zhou, Y.; Long, Z.; Wang, X.; Li, J.; Wang, J. Mesoporous Polyoxometalate-Based Ionic Hybrid as a Triphasic Catalyst for Oxidation of Benzyl Alcohol with H2O2 on Water. ACS Appl. Mater. Interfaces 2014, 6, 4438–4446. [Google Scholar] [CrossRef]
  61. Tamizhdurai, P.; Narayanan, S.; Kumaran, R.; Mangesh, V.; Kavitha, C.; Lakshmi, N.V.; Ragupathi, C.; Alothman, Z.A.; Ouladsmane, M.; Mani, G. Catalytic activity of ratio-dependent SBA-15 supported cerium/Pt catalysts for highly selective oxidation reaction of benzyl alcohol to benzaldehyde. Adv. Powder Technol. 2021, 32, 4286–4294. [Google Scholar] [CrossRef]
  62. Rezaei, A.; Mohammadi, Y.; Ramazani, A.; Zheng, H. Ultrasound-assisted pseudohomogeneous tungstate catalyst for selective oxidation of alcohols to aldehydes. Sci. Rep. 2022, 12, 3367. [Google Scholar] [CrossRef] [PubMed]
  63. Nomiya, K.; Yagishita, K.; Nemoto, Y.; Kamataki, T.-A. Functional action of Keggin-type mono-vanadiumc V-substituted heteropolymolybdate as a single species on catalytic hydroxylation of benzene in the presence of hydrogen peroxide. J. Mol. Catal. A Chem. 1997, 126, 43–53. [Google Scholar] [CrossRef]
  64. Ge, H.; Leng, Y.; Zhou, C.; Wang, J. Direct hydroxylation of benzene to phenol with molecular oxygen over phase transfer catalysts: Cyclodextrins complexes with vanadium-substituted heteropoly acids. Catal. Lett. 2008, 124, 324–329. [Google Scholar] [CrossRef]
Figure 1. Colors of the synthesized materials: (a) (TBA)3PW, (b) (TBA)4PWV, (c) (TBA)5PWV2.
Figure 1. Colors of the synthesized materials: (a) (TBA)3PW, (b) (TBA)4PWV, (c) (TBA)5PWV2.
Catalysts 12 00507 g001
Figure 2. FT-IR spectra of pure and V-substituted Mo (a) and W (b) Keggin-type polyoxometalates TBA salts.
Figure 2. FT-IR spectra of pure and V-substituted Mo (a) and W (b) Keggin-type polyoxometalates TBA salts.
Catalysts 12 00507 g002
Figure 3. XRD patterns of pure and V-substituted Mo (a) and W (b) Keggin-type polyoxometalates. The red symbol represents the comparison between equivalent diffractograms with JCPDS cards.
Figure 3. XRD patterns of pure and V-substituted Mo (a) and W (b) Keggin-type polyoxometalates. The red symbol represents the comparison between equivalent diffractograms with JCPDS cards.
Catalysts 12 00507 g003
Figure 4. DRS UV-Vis spectra of pure and V-substituted Mo Keggin-type polyoxometalates.
Figure 4. DRS UV-Vis spectra of pure and V-substituted Mo Keggin-type polyoxometalates.
Catalysts 12 00507 g004
Figure 5. N2-adsorption isotherm of pure and V-substituted Mo (a) and W (b) Keggin-type polyoxometalates.
Figure 5. N2-adsorption isotherm of pure and V-substituted Mo (a) and W (b) Keggin-type polyoxometalates.
Catalysts 12 00507 g005
Figure 6. The SEM images of the Mo-series synthesized materials: (a) (TBA)3PMo, (b) (TBA)4PMoV, (c) (TBA)5PMoV2.
Figure 6. The SEM images of the Mo-series synthesized materials: (a) (TBA)3PMo, (b) (TBA)4PMoV, (c) (TBA)5PMoV2.
Catalysts 12 00507 g006
Figure 7. HLPC-DAD chromatogram of oxidation products of benzyl alcohol at 4 h. Benzaldehyde was the only reaction product, as confirmed by GC-MS.
Figure 7. HLPC-DAD chromatogram of oxidation products of benzyl alcohol at 4 h. Benzaldehyde was the only reaction product, as confirmed by GC-MS.
Catalysts 12 00507 g007
Figure 8. Total conversions (a) (%, color bars) and rate constants (b) (s−1, grey bars) of the liquid-phase aerobic oxidation of benzyl alcohol at 4h for pure and substituted Mo and W catalysts (n = 3).
Figure 8. Total conversions (a) (%, color bars) and rate constants (b) (s−1, grey bars) of the liquid-phase aerobic oxidation of benzyl alcohol at 4h for pure and substituted Mo and W catalysts (n = 3).
Catalysts 12 00507 g008
Figure 9. Reuse of PMoV2 at 4 h, 5 h, 6 h, and 7 h. The black bar represents the first cycle, the dark-gray bar the second, and finally, the light-gray bar the third, for each time.
Figure 9. Reuse of PMoV2 at 4 h, 5 h, 6 h, and 7 h. The black bar represents the first cycle, the dark-gray bar the second, and finally, the light-gray bar the third, for each time.
Catalysts 12 00507 g009
Figure 10. FT-IR (a) and XRD (b) spectra of reused (red line) and fresh (green line) (TBA)5PMoV2 catalyst; (c) SEM micrography of the reused (TBA)5PMoV2 catalyst.
Figure 10. FT-IR (a) and XRD (b) spectra of reused (red line) and fresh (green line) (TBA)5PMoV2 catalyst; (c) SEM micrography of the reused (TBA)5PMoV2 catalyst.
Catalysts 12 00507 g010
Table 1. Absorption edge energy of the pure and substituted Mo and W catalysts.
Table 1. Absorption edge energy of the pure and substituted Mo and W catalysts.
CatalystAbsorption Edge
(nm)
(TBA)3PMo495
(TBA)4PMoV544
(TBA)5PMoV2569
(TBA)3PW356
(TBA)4PWV493
(TBA)5PWV2597
Table 2. Textural properties of the pure and substituted Mo and W catalysts.
Table 2. Textural properties of the pure and substituted Mo and W catalysts.
CatalystSBET
(m²/g)
(TBA)3PMo3
(TBA)4PMoV9
(TBA)5PMoV214
(TBA)3PW5
(TBA)4PWV8
(TBA)5PWV2 14
Table 3. Catalytic performance in the oxidation of benzyl alcohol for various catalysts.
Table 3. Catalytic performance in the oxidation of benzyl alcohol for various catalysts.
CatalystSolventOxidantTime (h)Temperature (°C)Conversion (%)Selectivity a (%)Ref.
Au-PdMeOHH2O20.55011.3>85[57]
Pd/CaSUP-H2O2 (30%)88088.089[58]
NiOx-CuOx/SBA-15Organic-59088.976[59]
[TMGHA]H0.6PW2.4H2OH2O2 (30%)69097.993[60]
Ce-Pt/SBA-15AcNTBHP79098.899[61]
A-CDQs/WH2OH2O20.052598.093[62]
(TBA)5PMoV2MeOH:H2OO2417093>99This work
a Selectivity towards benzaldehyde.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Díaz, J.; Pizzio, L.R.; Pecchi, G.; Campos, C.H.; Azócar, L.; Briones, R.; Romero, R.; Henríquez, A.; Gaigneaux, E.M.; Contreras, D. Tetrabutyl Ammonium Salts of Keggin-Type Vanadium-Substituted Phosphomolybdates and Phosphotungstates for Selective Aerobic Catalytic Oxidation of Benzyl Alcohol. Catalysts 2022, 12, 507. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12050507

AMA Style

Díaz J, Pizzio LR, Pecchi G, Campos CH, Azócar L, Briones R, Romero R, Henríquez A, Gaigneaux EM, Contreras D. Tetrabutyl Ammonium Salts of Keggin-Type Vanadium-Substituted Phosphomolybdates and Phosphotungstates for Selective Aerobic Catalytic Oxidation of Benzyl Alcohol. Catalysts. 2022; 12(5):507. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12050507

Chicago/Turabian Style

Díaz, Juan, Luis R. Pizzio, Gina Pecchi, Cristian H. Campos, Laura Azócar, Rodrigo Briones, Romina Romero, Adolfo Henríquez, Eric M. Gaigneaux, and David Contreras. 2022. "Tetrabutyl Ammonium Salts of Keggin-Type Vanadium-Substituted Phosphomolybdates and Phosphotungstates for Selective Aerobic Catalytic Oxidation of Benzyl Alcohol" Catalysts 12, no. 5: 507. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12050507

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop