Next Article in Journal
Effect of Grain Size and Surface Roughness on the Normal Coefficient of Restitution of Single Grains
Next Article in Special Issue
Molybdenum-, Vanadium-, and Tungsten-Containing Materials for Catalytic Applications
Previous Article in Journal
Chitosan-TiO2: A Versatile Hybrid Composite
Previous Article in Special Issue
H3PMo12O40 Immobilized on Amine Functionalized SBA-15 as a Catalyst for Aldose Epimerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vanadium-Substituted Phosphomolybdic Acids for the Aerobic Cleavage of Lignin Models—Mechanistic Aspect and Extension to Lignin

1
Centre National de la Recherche Scientifique, UMR 7197, Laboratoire de Réactivité de Surface (LRS), Sorbonne Université, F-75005 Paris, France
2
Centre National de la Recherche Scientifique, UMR 7190, Institut Jean le Rond d’Alembert, Sorbonne Université, F-75005 Paris, France
*
Author to whom correspondence should be addressed.
Materials 2020, 13(4), 812; https://doi.org/10.3390/ma13040812
Submission received: 23 January 2020 / Revised: 7 February 2020 / Accepted: 7 February 2020 / Published: 11 February 2020

Abstract

:
This work deals with the aerobic oxidative cleavage of C-C and C-O bonds catalyzed by the Keggin-type phosphovanadomolybdic acid (H6[PMo9V3O40], noted H6PV3). The latter was synthesized by an adapted hydrothermal procedure classically used for lower vanadium content and was tested as a catalyst for the aerobic cleavage of 2-phenoxyacetophenone (noted K1HH) and 1-phenyl-2-phenoxyethanol (A1HH) used as two lignin models. The operative conditions (solvent, catalytic loading, etc.) were adjusted on K1HH and extrapolated to A1HH. The cleavage of the alcohol model required more drastic conditions and therefore further optimization. Preliminary attempts on an Organosolv wheat straw lignin were performed too. From the kinetic study, high performance liquid chromatography (HPLC) and gas chromatography–mass spectrometry (GC-MS) data, a mechanism of the cleavage of both models was proposed.

Graphical Abstract

1. Introduction

Lignin is the second most abundant natural polymer on our planet and the only natural, renewable and abundant source of aromatic compounds. It has raised increasing interest due to the depletion of oil, from which aromatic compounds are currently mainly produced. One of the most outstanding lignin valorization routes is its oxidation to aromatic platform molecules such as vanillin obtained from C-C cleavage [1]. Aerobic lignin cleavage in alkali media was already set up at the industrial scale on the 1930s. The major problem was the elimination of caustic wastes conjugated with a rather low yield of vanillin [2]. So, such of process is currently not competitive against the oil-based routes for vanillin production and has to be improved. In the present paper, O2 will be used as it is the most sustainable oxidant. However, it needs to be activated by a catalyst. Several metal-containing catalysts were assessed for the aerobic oxidative cleavage of lignin [3,4]. One of the most used metals is vanadium [5,6,7]. Keggin-type phosphovanadomolybdic acids (H3+xPMo12−xVxO40, nH2O) noted H3+xPVx [8,9,10] are among the vanadium-based catalysts reported. For two decades, since the contributions of Weinstock et al. [11] and Evtuguin et al. [12], Keggin-type H3+xPVx are well-known to be adequate for wood pulp oxygen delignification and more largely for aerobic oxidation [13]. Also, these catalysts already proved to be efficient for the C-C cleavage of α-ketols (room temperature) such as benzoin, a lignin model or cyclohexanone (70 °C) in an atmosphere of molecular oxygen [14,15,16].
Since lignin is an extremely complex molecule, preliminary studies of its oxidative cleavage are usually carried out on models. Several linkages such as β-5, β-β and 5-5 are represented in lignin, but the β-O-4 bond (Figure 1) is always the major one in its native form [16]. As a result, many V-catalysts [17,18,19,20,21,22,23,24,25,26,27] were assessed for the aerobic oxidative cleavage of models bearing these linkages.
In the present contribution, the considered models are, on one hand 1-phenyl-2-phenoxyethanol noted A1HH (already tested by Hanson et al. in presence of V(+V)(O)(dipic)(OiPr), (dipic = dipicolinate)) and, on the other hand the ketone, 2-phenoxyacetophenone, K1HH, an oxidation product from the alcohol model. To the best of our knowledge, H3+xPVx catalysts were never tested on such ketone models before. The active sites of H3+xPVx are rather the vanadium addenda and it is known that the oxidative power of H3+xPVx increases with vanadium content (mainly because of the reversible generation of pervanadyl cations VO2+ which are stronger oxidant than their parent H3+xPVx) [28,29]. However, an increase of vanadium degrades the stability of the H3+xPVx [30] that may lead to loss of activity as observed by Odyakov et al. As a result, in this study, H6PV3 will be chosen as in the work of El Aakel et al. [14], because it seems to be the best compromise between oxidative power and stability. H6PV3 was synthesized through the hydrothermal pathway which just consists in refluxing the starting oxides and phosphoric acid in water without any addition of an external reagent [31] on the contrary to the most used etherate [32] (that involves a hazardous solvent [33] and strong acids) and oxo-peroxo [34,35] procedures usually used for such material. Already, such a pathway was attempted by Grate et al. for the synthesis of a H6PV3 0.30 M in water. Even after seven days of reflux, the inclusion of vanadium was incomplete [36]. Therefore, in the present case, the procedure will be carried out in more diluted conditions (0.17 M) to favor the hydrothermal attack of V2O5.

2. Materials and Methods

2.1. Materials Synthesis

H6PV3 was synthesized through the attack of MoO3 (Acros Organics, Illkirch, France) and V2O5 (Sigma-Aldrich Chemie S.a.r.l., Saint-Quentin Fallavier, France) by H3PO4 (purity 85%, Carlo Erba, Val-de-Reuil, France) in refluxing water according to a procedure adapted from Kern et al. [31]. Hence, 0.39 g (3.3 mmol) of H3PO4 was dissolved in 200 mL of water and 3.9 g of MoO3 (27.3 mmol) and 0.92 g (5.1 mmol) of V2O5 were added.

2.2. H6PV3 Characterization

X-ray powder diffraction analyses (XRD, Advance D8, Bruker, Champs-sur-Marne, France) were performed using a Cu Kα radiation source without monochromator. The diffraction patterns were collected from 5° to 50°, with a scanning rate of 0.34° min−1. The crystalline phases were identified through the Rietveld analysis of the diffraction patterns (factors RP and Chi2), using the Fullprof® Suite software (program Fullprof. 2k, version 5.40, CEA, Saclay, France). Liquid 31P nuclear magnetic resonance analyses were performed on a 400 MHz apparatus (NMR, Bruker, Wissembourg, France). Hence, 30 mg of the solid was dissolved in 250 µL of D2O (Eurisotop) and 250 µL of H2O, then 7.5 µL of dioxane (SDS, Carlo-Erba, Val-de-Reuil, France) were added to the solution. For each analysis, 16 scans were recorded with a relaxation delay of 32 s. Inductive coupled plasma analyses (ICP) were carried out by Crealins (Lyon, Villeurbanne France) on an ICP Thermo-Fischer iCAM. ICP analysis allowed the determination of the value of x. Thermogravimetric analyses (TGA, Cp SDTQ600 system, TA Instruments, New Castle, USA) were performed in order to determine the hydration index of H3+xPVx, noted n. The temperature was increased to 600 °C at a heating rate of 10 °C/min under an air flow of 100 mL min−1.

2.3. Catalytic Tests

The efficiency of H6PV3 for the aerobic cleavage in atmospheric conditions was tested on K1HH (0.32 g, 1.5 mmol), A1HH (0.32 g, 1.5 mmol) or the acetate derivative of A1HH, EstαA1HH (0.41 g, 1.5 mmol). The preparation procedure of all these compounds is detailed in the Supporting Information. The substrate, and then H6PV3 (0–88 mg), were dissolved in 15 mL of the solvent in a Schlenk tube connected to a gas burette system to monitor the molecular oxygen consumption. The Schlenk tube was purged with molecular oxygen and then heated during 24 h.
Also, some tests on A1HH were carried out at higher pressure. In a typical experiment, 0.54 g of the substrate (2.5 mmol) and 61.4 mg of H6PV3 (Mo + V 15 mol%) were dissolved in 25 mL of solvent (MeCN + 10 vol% AcOH) in a Teflon container that was further placed in a 100 mL autoclave. The mixture became instantaneously green due to the reduction of the catalyst. Once the autoclave was sealed, the reaction medium was purged three times by molecular oxygen at room temperature. The pressure was set to 5 bar and then heated at 80 °C or 120 °C under stirring. Once the reaction was completed, the recovered solution is orange-brown.

2.4. Monitoring of the Tests

After cooling the reaction mixture, 0.5 mL was sampled and dropped in a 20 mL volumetric flask and diluted by H2O-MeOH-AcOH 49:49:2 for high performance liquid chromatography (HPLC) analysis (LC-20AD, Shimadzu France, Marne-la-Vallée, France) using a Shim-pack column (4.6 × 100 mm, 2.2 µm) heated to 40 °C in a CT0-10AS oven. The mobile phase was constituted of aqueous AcOH (0.5 vol%) and methanol (HPLC grade, VWR Chemicals) used at a flow rate of 0.4 mL min−1 (0–13 min: methanol 40%, 13–17 min: gradient from 40 to 60% of methanol, 17–33 min: methanol 60%, 33–37 min: gradient from 60 to 40% of methanol and 37–50 min: methanol 40%). The compounds were detected by an SPD-M20A UV detector (210, 220 and 254 nm (Figure S1)). Benzaldehyde, benzoic acid and phenols were quantified according calibration curves (see Supporting Information, Figure S2). A typical example is given on the Supporting Information (Figure S1). Benzaldehyde, benzoic acid and phenol were quantified according calibration curves (see Supporting Information, Figure S2 and Table S1). The other products (of the C-O cleavage such as phenylglyoxylic acid as well as quinones and so forth, detected by HPLC (see Figure S1)) were not taken into account.
The reaction medium was also analyzed without dilution by gas chromatography–mass spectrometry (GC-MS) in duplicate on a GC-2010 Plus gas chromatograph using a HP-5MS ((5%-phenyl)-methylpolysiloxane) column. The temperature of the injector was set to 250 °C. The oven temperature raised from 70 °C to 250 °C with a ramp of 5 °C min−1 and the temperature was held at 250°C during 10 min. Helium (1.5 mL min−1) was used as the carrier gas. The products were detected by a GCMS-QP2010 SE mass spectrometer with a time delay of 4 min (Table S2). The ion source and the mass detector were heated to 200 and 250°C respectively and the voltage was 0.2 kV. The mass spectra were recorded for 35 < m/z < 300 every 0.3 s.

2.5. Preliminary Tests on an Organosolv Lignin

These tests were carried out on a purified wheat straw Organosolv lignin (extracted from wheat straw according to a patent FR 2926824 [37] and purified according to Mbotchak et al. [38]) in the presence of H6PV3 in the conditions of A1HH cleavage. The targeted products were p-hydroxybenzaldehyde, vanillin, syringaldehyde and their corresponding acids. Experiments were carried out directly with 5 bars of molecular oxygen at 120 °C. Hence, 0.85 g of lignin and 59.6 mg of H6PV3 (Mo + V 15 mol% according to the substrate) were weighted and contacted with 25 mL of MeCN − 10 vol% AcOH in an autoclave. The procedure used was then similar to that of A1HH oxidation under pressure. At the end of the reaction, since lignin is barely soluble, the mixture had to be filtered on a sintered filter (porosity n 4). An aliquot of the filtrate (20 µL) containing the targeted phenolic aldehydes and acids was poured in a test tube capped by a septum in order to evaporate the solvent. Then, 20 µL of pyridine, 10 µL of a solution of an internal standard (biphenyl, 10 mM in acetonitrile) and 250 µL of BSTFA (N,O-Bis(trimethylsilyl)trifluoroacetamide) were added and the mixture was stirred overnight. Afterwards, the obtained solution was analyzed by GC-MS in duplicate.

3. Results

H6PV3 was characterized by XRD and liquid 31P NMR to validate the Keggin-type structure and to obtain its yield. Then, the number of vanadium equivalents and the hydration index noted x and n were determined by ICP and thermogravimetric analysis (TGA) respectively. H6PV3 was then tested for K1HH cleavage. Meanwhile, the operative conditions were optimized. Then, H6PV3 was assessed in optimized conditions for A1HH and then an Organosolv wheat straw lignin.

3.1. H6PV3 Synthetic Procedure

H6PV3 synthesized using the hydrothermal pathway was characterized by XRD and 31P NMR. The X-Ray profile of H6PV3, as well as of its starting oxides acquired in the same conditions, are given on Figure 2.
The diffractogram of H6PV3 was shown to be very similar to that of tridecahydrated phosphomolybdic acid, H3PMo12O40, 13 H2O (JCPDS 00-043-0317, see Figure S3) which adopts a triclinic ( P 1 ¯ ) structure. The main characteristic peaks of this Keggin-type structure are expected at c.a. 9° and c.a. 28°. It has to be noted that the peaks at c.a. 9° are not present in the profiles of the starting oxides and are clearly related to the triclinic structure of V3. However, H6PV3 turned out to be constituted of at least two crystalline phases since RP = 15.1% and Chi2 = 13.1% (Table 1). The second crystalline phase could be unreacted vanadium oxide or a H6PV3 having a different hydration index. The cell parameters of the main phase of H6PV3 were calculated on Fullprof® and are given in Table 1 (an extract of Fullprof is given on Figure S4).
Liquid 31P NMR (Figure 3) was performed in order to establish the composition of H6PV3 in solution. To do this, the structure of H6PV3 had to be stabilized by dioxane [39] at a final pH of about 1.
The spectrum of H6PV3 was more complex compared to those of H3+xPVx having lower amounts of V addenda [36,40]. On one hand, 13 isomers (taking account only the α isomers) can be envisaged and, on the other hand, the introduction of vanadium leads to a weaker acidity strength [40]. As a result, the additional peaks from −2.9 to −3.25 ppm observed for H6PV3 could be associated to anionic and protonated forms of H6PV3 (pKa ≈ 3). Also, pervanadyl cations (VO2+) are known to be more easily ejected from the Keggin-type sphere leading to lacunar H3+xPVx among others [28]. NMR enabled to confirm the Keggin-type structure. As shown on Figures S5 and S6, the maximal relaxation time was established to 1.02 ± 0.11 s. So, the relaxation time set in each acquisition was higher than 5 times the relaxation delay of every phosphorous nucleus. As a result, the integration was considered to be quantitative and, because reactants were added in stoichiometric amounts, the area of the peak at 0 ppm enabled to evaluate the amount of unreacted H3PO4 and so the H3+xPVx yield, assuming that H3PO4 was the only by-product. The value, hence calculated, was 79%.
Once the Keggin-type structure validated by XRD and confirmed by liquid 31P NMR, the number of vanadium equivalents in H6PV3 solid was then calculated from the weight fraction of molybdenum (46.16 wt%) and vanadium (7.06 wt%) obtained by ICP. It was shown to be equal to 2.680 ± 0.005 (vs a theoretical × of 3). The hydration index n calculated by thermogravimetric analysis and differential scanning calorimetry (TGA-DSC was 13 (Figure 4) through the first weight loss [41], which is in good agreement with the XRD data. So, the general formula for H6PV3 is H5.68PMo9.32V2.68O40, 13H2O. The second weight loss weight loss may enable to calculate the number of vanadium equivalents, but, here, the loss of hydration and constitutive water are overlapped [42].

3.2. K1HH aerobic cleavage

Firstly, H6PV3 was assessed for the aerobic cleavage of K1HH under atmospheric pressure of oxygen in pure solvents. The solution was heated to the targeted temperature during 24 h. In such conditions, phenol (PhOH), benzaldehyde (PhCHO) and benzoic acid (PhCOOH) were obtained as the main products (Figure 5).
The amount of p-quinone could not be determined accurately but the latter was rather small. The conversion of K1HH, the yields of the main products as well as the carbon balance values (CB) are given in Table 2. The products from both C-C (PhCHO and PhCOOH) and C-O (PhOH) cleavage and the starting material were taken into account in the calculation of the carbon balance. As usually observed, the carbon balance tends to decrease with increasing conversion as the result of the occurrence of side reactions. It has to be noted here that, the maximum value of the carbon mass is normally lower than 100%. Indeed, C1 products such as formaldehyde, formic acid, CO2 and so forth, coming from the transformation of the CH2 group of K1HH could not be detected quantitatively and were not taken into account. Assuming all of this, CB should not take values higher than C B m a x % = 1400 C o n v   % 14 .
Several solvents were tested. Refluxing conditions (excepted for the test in pure acetic acid) in presence of H6PV3 (Mo + V 15 mol%) during 24 h were applied. As a starting point, methanol (Table 2, entry 1) and acetic acid (entry 3) were used like El Aakel et al. did [18], for the C-C bond cleavage of cyclic α-hydroxyketones in presence of V3. Acetonitrile was assessed too (Table 2, entry 2). The conversion of K1HH in refluxing acetonitrile (Table 2, entry 2) was not so high (48%, brown) and turned out to be lower in refluxing methanol (14%, green, entry 1). The highest conversion of K1HH and yields of PhOH and PhCOOH were obtained in acetic acid (Table 2, entry 3). However, lignin is rich of alcohol moieties that may be esterified by acetic acid, giving rise to a lignin more recalcitrant to oxidation [43]. Consequently, in spite of these results, acetonitrile was preferred to acetic acid provided that some further optimization would be done.
Mixtures of acetonitrile and acetic acid with varying AcOH concentrations (0 ≤ z ≤ 20% AcOH in volume) have also been tested as solvents at 82 °C (Table 2, entries 4–8) [44,45,46]. The conversion of K1HH as well as the yields of phenol coming from C-O and of benzoic acid from C-C cleavage, respectively raised with the acetic acid content. A z value of 10 vol. % was found to be an optimum. Indeed, at 20 vol. % (Table 2, entry 8 vs 6), the carbon balance decreased (72% vs 80%). Since C-O bond is weaker than C-C bond, its cleavage was favored giving rise to an enhanced phenol production compared to benzaldehyde and benzoic acid. Some minor products could be detected by HPLC like phenylglyoxylic acid for z = 2.5 (yield 2.7%, estimated by HPLC) and z = 5 (yield 4.5%). Furthermore, benzaldehyde yield was found to be always lower than those of phenol and benzoic acid. The maximum yield of benzaldehyde was reached for z = 5 (Table 2, entry 5) and the PhCHO/PhCOOH ratio was maximized for z = 2.5 (Table 2, entry 4) and tends to decrease for a higher content of acetic acid until the disappearance of benzaldehyde in pure acetic acid (Table 2, entry 3). So, acetic acid appears to favor the benzaldehyde over-oxidation.
The test in MeCN − 10 vol% AcOH was repeated at 65 °C (Table 2, entry 7). From the comparison with the test in methanol (Table 2, entry 1), the test in acidified acetonitrile gave rise to better conversion and yields of cleavage products. The (PhCHO + PhCOOH)/PhOH ratio was equal to 1.2 at 65 °C and 1 at 80 °C.
A series of experiments were performed in order to evaluate the influence of the H6PV3 loading (from 7 to 36 mol% of Mo + V, Table 2, entries 6, 9–11). Logically, the higher the metal loading, the higher the conversion of the substrate. At low Mo + V loading (Table 2, entry 10, 7 mol%), the conversion was rather low (43%). At the highest metal loading (Table 2, entry 11, 36 mol%), the conversion was expectedly higher but the carbon balance tended to be rather low (58% only, compared to 80% for 15 mol% Mo + V). For these reasons, a metal loading of 15 mol% was found to be a good compromise. Besides, in the case of the tests performed at high metal loading (Table 2, entry 11), the oxygen consumption per mole of converted substrate was slightly lower probably because the catalyst might behave as an oxygen donor [47]. Materials such as H6PV3 are characterized by rather strong acidic properties, that is the reason why we have decided to check also the influence of the acidity on the cleavage of K1HH. This was done owing to a reference test (no metal) performed with H2SO4 (Table 2, entry 9) which showed that the conversion of K1HH and especially the yields of PhCHO and PhCOOH were lower than with V3. Phenol was by far the main product meaning that the C-C cleavage requires the presence of a metallic catalyst whereas C-O linkage may be cleaved more easily, just by acidity.
Most of the previous tests in MeCN − 10 vol% AcOH were performed at 80 °C within 24 h. Here, we propose to monitor, during 48 h, the K1HH conversion, the carbon balance as well as the yields of the three main products in our reference conditions, that is, K1HH 100 mM, H6PV3 (Mo + V 15 mol%), MeCN − 10 vol% AcOH (15 mL), T = 80 °C (see Figure 6). The yield of p-quinone was also calculated, though with a low precision. As expected, the conversion of K1HH raised whereas the carbon balance decreased throughout the reaction. The yield of PhCHO increased until 24 h with a maximum of 12%, then it slowly decreased to 9%. PhCOOH production reached also its maximum (46%) within 37 h. After, it did not change significantly. First, before 24 h, the reaction duration was positive for PhOH formation with a maximum of 67%. Then, due to over-oxidation into p-quinone as observed after 24 h, the yield of PhOH decreased until 46% till 37 h. This led to a significant decrease of the carbon balance. For short reaction times (<24 h), the (PhCHO + PhCOOH)/PhOH ratio was often lower than 1 meaning that PhOH is produced in greater extent than the two other main products. As a result, it is highly probable that part of phenol was issued from C-O cleavage only. The detection by HPLC of C8-products such as phenylglyoxylic acid confirms such hypothesis. The yield of the latter could be estimated to c.a. 6% showing no significant evolution between 37 and 48 h. Hence, the quinone production was limited and took place to a much smaller extent compared to phenol formation from K1HH. Furthermore, no molecular oxygen volume variation was observed between 37 and 48 h while the conversion of K1HH still raised in the meantime. Otherwise, the molecular oxygen consumption per mole of converted substrate was quite constant with a value of 1.5 mol of O2 per mole of consumed K1HH.
Besides, the quantity of gaseous molecular oxygen is much higher than dissolved molecular oxygen. So, it can be hypothesized that dissolved oxygen is constantly renewed and so, the concentration of molecular oxygen is constant (order degeneration). We tested a general expression d K 1 HH dt = k app , K K 1 HH q K with qK order values different from 1 and it appeared that the best modeling of the experimental data corresponds to [K1HH]−0.25 = f(t) (Figure S7) curve, that is, (after derivation) qK = 1.25, which is close to 1.

3.3. A1HH aerobic Cleavage

K1HH is not the best model of lignin. A1HH has a structure closer to lignin and will be now considered as the substrate (Figure 7).
A1HH was tested first in the conditions optimized for K1HH, that is, in the presence of H6PV3 (Mo + V 15 mol%) in MeCN + 10 vol% AcOH under atmospheric pressure of O2 (Table 3, entries 1–2).
Undoubtedly, A1HH proved to be more resistant than K1HH. Under atmospheric pressure, the oxidation of the alcohol function was favored as well as the formation of the esterification (by acetic acid) and dehydration products, noted EstαA1HH and D1HH, respectively. In the presence of H2SO4, instead of H6PV3 (Table 3, entry 2), the conversion fell to 9.1% (vs 35% in presence of V3, Table 3, entry 1), the yield of phenol was 1.9% (vs 4.6%) and benzoic acid was not detected. Also, the ester was tested in the same conditions in the presence of H6PV3. A conversion of 7% and only traces of cleavage products were obtained meaning that the esterification is a dead-end reaction. Therefore, such a reaction has to be mitigated.
The cleavage reaction of A1HH was also monitored for 48 h (Figure 8). During that time period, the conversion of A1HH raised until 67% (instead of 35% within 24 h). Clearly, the oxidation of A1HH was slow compared to that of K1HH under the same conditions. The yields of phenol and benzoic acid raised both until 10% and benzaldehyde yield was stable from 8 to 24 h meaning that its production and its oxidation took place at very similar rates. K1HH seemed to accumulate (the yield of K1HH raised until 17% at 24 h, then the latter was converted since its yield was 11% after 48 h showing that, in presence of A1HH, H6PV3 becomes less efficient for the aerobic oxidative cleavage of K1HH. The side products EstαA1HH (calibrated at 254 nm, see Figure S2) and D1HH (identified by 1H NMR) accumulated during the course of the reaction. With increasing time, the phenol selectivity was stable whereas the selectivity in C-C cleavage increased. The carbon balance decreased after 37 h meaning that side reactions were also accelerated.
Similarly to K1HH cleavage, the rate law for A1HH can be written d A 1 HH dt =   k app , A A 1 HH q A . Since the best fit with the experimental data (max. R² = 0.9999) was obtained with [A1HH]−0.7 = f(t) (Figure S7), this gave, after derivation, qA = 1.7. The values of kapp,A (0.0009 M−0.25 s−1) and kapp,K (0.0036 M−0.7 s−1) were also determined (Figure 8a–b and Figure S7). From these results, a slower kinetics of A1HH (vs K1HH) conversion and a higher impact of the substrate concentration in the case of A1HH could be evidenced.
More stringent conditions had to be implemented in order to favor the formation of oxidative cleavage products. In a first attempt, the O2 pressure was raised from 1 to 5 bar without modifying the temperature (80 °C; Table 3, entry 3). Such rise of O2 pressure favored the conversion (54 vs 35%, Table 3, entry 3 vs 1) as well as the yield of PhCHO and PhCOOH, that is, the selectivity for C-C cleavage products. Nonetheless, dehydration and esterification took place at a higher extent. Since then, the duration of the test had to be shortened to 6 h (entry 4 vs 3) instead of 24 h and A1HH conversion decreased, as expected, from 54 to 33%.
Another set of conditions was tested. Hence, the temperature was raised to 120 °C and the duration was shortened to 2 h (Table 3, entry 5) to preserve the carbon balance. After 2 h, the conversion was 44% (vs 33% at 80°C after 6 h, Table 3, entry 4). The temperature increase turned out to have a positive impact (Table 3, entry 5 vs 4) on the selectivity for the C-O (16% vs 11%) and C-C cleavage (32% vs 19%). Moreover, EstαA1HH (11% vs 9.9%) was produced at similar extent and the carbon balance was not significantly degraded (96% vs 99%). Therefore, another test was performed at 120 °C during 6 h (entry 6). Compared to 120 °C/2 h (Table 3, entry 5), the conversion was expectedly improved (90% vs 44%) and the yield of benzoic acid was higher (19% vs 7.5%). However, the yield of phenol was not improved significantly and the carbon balance fell to 55% (vs 96%) meaning there is a higher carbon loss due to side reactions.
To conclude, ideally A1HH cleavage tests should be performed at 120 °C during 2 h (Table 3, entry 5) since a reasonable compromise can be reached between the A1HH conversion, the yields of PhOH, PhCHO and PhCOOH and a correct mass balance in these conditions.

3.4. Lignin Aerobic Cleavage: Preliminary Tests

The catalytic tests on lignin (WSLp) were performed directly under 5 bars of O2 (measured at R.T.) at 120° during either 2 or 6 h (Figure 9) like in the case of A1HH.
The expected aromatic aldehydes noted H, V and S could be detected by GC-MS (see Table S3). However, the expected phenolic acids were barely evidenced by GC-MS. Since those acids are difficult to detect with apolar GC columns and are not very stable at high temperature [48], the reaction mixture had to be treated overnight by an excess of N,O-Bis(trimethylsilyl)trifluoroacetamide (BSTFA) for quantification of the trimethylsilyl derivatives by GC-MS. Alcohols and acids were both derivatized. The reaction medium was also injected without silylation to check the presence of by-products such as 2,6-DiMethoxyBenzoQuinone (2,6-DMBQ) or acetosyringone. The retention times of the main derivatized oxidation products identified are given in Table S3. The GC-MS analyses of the reaction media before silylation also emphasized the formation of p-quinones derivatives such as 2,6-DMBQ which are products of the C(Ar)-C cleavage, as observed by Evtuguin et al. [27]. Such a reaction would take place due to the presence of phenolic functions since these quinones were observed only as minor by-product in the case of the oxidative cleavage of the non-phenolic models, especially for A1HH. Acetovanillone was produced too. This may be originated from either an oxidative or a non-oxidative C-O cleavage, accelerated by the acidity brought by the catalyst [25,27].
The yields of the targeted soluble products, determined by GC-MS (calibration curves given on Figure S8) were rather low with a sum for H, HA, V, VA, S and SA (expressed in mol%) lower than 3% in the best case (120 °C, 6 h, Table 4, entry 3). The catalyst turned out to be necessary for lignin aerobic cleavage (entries 1 vs. 2) since V, VA and HA monomers were not detected after 2 h at 120 °C in the absence of V3. Moreover, the extent of lignin dissolution (9–10 wt%) was rather unchanged by the catalyst. At 120 °C, a longer reaction time had a positive impact on the yield of cleavage products (Table 4, entries 3 vs. 1).

4. Discussion

4.1. H6PV3 Synthesis

The targeted x value (3) could be almost obtained by following the hydrothermal synthesis (experimental ×: 2.7). Yet, compared to Reference 31 (applied for V1 synthesis), a longer attack (6.5 h instead of 3 h) and more diluted conditions were needed to dissolve the oxides. However, part of the oxides did not dissolve and had to be eliminated by filtration despite the use of more diluted conditions and long heating time. The solid residue was found to be richer in vanadium (wMo = 23 wt.%, wv = 38 wt.%) than the initial mixture of oxides. A reflux of 24 h was tested to enhance the dissolution but the amount of residual oxide did not decrease significantly meaning that a longer attack is not efficient. Yet, inclusion of vanadium was better compared to that reported by GRATE et al. [36] (V2.4 instead of V3) despite a much longer heating step (6.5 h vs 168 h). It is known that H6PV3 has a rather low thermal instability and stability to the hydrolysis of VO2+ generated from the parent H6PV3 [30]. As a result, the long heating duration applied by Grate et al. may have favored H6PV3 destruction rather than the dissolution of the starting oxides.

4.2. Evolution of the Catalyst

After K1HH oxidation (Table 2, entry 6), the catalyst was recovered and analyzed by 31P NMR showing only one peak at −4.2 ppm (Figure 10) instead of several ones as shown in Figure 3. The latter, measured at pH = 1, could be attributed to the phosphomolybdate anion. Having checked, in parallel, that the Keggin-type structure was preserved when only heating H6PV3 in MeCN - 10 vol% AcOH (80 °C), we may reasonably suggest that H6PV3 was not stable upon the reaction conditions, being transformed into a mixture of phosphomolybdate anion and, a priori, vanadyl (VO2+) or pervanadyl (VO2+) counter-cations. Indeed, it has been reported by Neumann and Khenkin that H3+xPVx compounds may lose a vanadyl cation when they are reduced by the substrate (here, H-Su) as the result of the increased distance between the heteroatom P and vanadium [49]. Such process is summarized in Equation (1) for x = 3.
PMo 9 V + V 3 O 40 6 + H Su = PMo 9 V + V 2 O 39 9 + V + I V O 2 + + H + + Su . .
Sometimes, H3+xPVx are also considered as the precursors of in situ-formed pervanadyl cations that would be the true catalysts as suggested by Bregeault et al. [50] and El Amrani et al. [51], characterized by a higher redox potential than parent H3+xPVx [29] and therefore may be more effective for oxidation, including aerobic cleavage. Such ejection is accompanied by the formation of an unstable lacunary polyoxometalate (POM) that generates a phosphomolybdate anion. Such reaction, favored by the acidity can be described by the equation Equation (2) [28] for x = 3:
H 6 PMo 9 V + V 3 O 40 + H + = H 5 PMo 9 V + V 2 O 37 + V + V O 2 + + H 2 O

4.3. K1HH and A1HH Cleavage

To get insight on the role of acetic acid, the O2 solubility in binary acetonitrile-acetic acid was also estimated using the equations established by Sato et al. [44] that relate the solubility of gases in solvents mixtures to their solubility in the corresponding pure solvents using Hansen parameters (details in the Supporting Information).
By analogy with El Aakel et al. [14], acetic acid and methanol were tested as solvents for the C-C cleavage of K1HH. In their work, methanol proved to be the most adequate solvent in the case of benzoin (Ph-CO-CH(OH)-Ph, r. T., atm. O2), which is not the case here for K1HH. Likewise, benzoin (H6PV3, r. T.) [14] or cyclohexanone (V2, 70 °C) [52], the active site in K1HH would be Cβ (see Figure 1). However, the C-H activation of Cβ would be easier in K1HH than in cyclohexanone as a result of the presence of an electro-donating phenoxy group stabilizing the transient radical.
We hypothesize that the enhancement of the yields of cleavage products in presence of acetic acid is related to a better solubility of O2. Indeed, with trifluoroacetic acid (TFA, 3.6 vol.%), which is a much stronger acid than acetic acid [53] but which has a similar affinity to molecular oxygen (solubility in TFA: 8.1 mM [54] vs 7.6 mM in acetic acid, see Table 2), no improvement of the C-C cleavage products was observed. The combined yield of PhCHO and PhCOOH was 38% vs 63% in presence of AcOH 10 vol.% meaning that acidity has a lower impact than O2 solubility. However, it can be noticed from Figure S9, that the addition of 10 vol.% of acetic acid in acetonitrile leads to a raise of molecular oxygen solubility of 11% only which means that acetic acid would have another role. Indeed, we suggest that, in presence of acetic acid, the C-C cleavage would be accelerated by a nucleophilic substitution reaction between the phenoxy group of K1HH and acetic acid leading to 2-acetoxyacetophenone observed by GC-MS. Indeed, such reaction can take place even without H6PV3 (Table 2, entry 9) in presence of 10 vol.% of acetic acid as shown by the large excess of phenol. The substitution reaction of -OPh by -OAc may also make the C-C oxidative cleavage easier due to a decrease of the steric hindrance. A last role of acetic acid could originate from its known ability to stabilize the pervanadyl cation VO2+ as proposed by MA et al. [23] in their work dealing with the aerobic oxidation of A1HH catalyzed by VO(acac)2 in MeOH-AcOH 8 vol%. Indeed, acetic acid would contribute to an easier regeneration of the active form of the catalyst through aerobic oxidation. However, the pH of the reaction media in our work should be lower compared to that measured by Ma et al. since H6PV3 and H3PMo12, obtained by the decomposition of H6PV3, are strong acids. Acetic acid being actually a weak acid in acetonitrile [53], the presence of H3+xPVx or of H3PMo12 should inhibit acetic acid deprotonation.
The present contribution dealt with the cleavage of two non-phenolic models K1HH and A1HH. Our results with the latter can be compared to those of Evtuguin et al. [27] who used a H8PV5 catalyst (480 mol% of Mo + V, O2 5 bar, acetone-water 70/30 v/v, 90 °C, 20 min) in the presence of another alcohol with electro-donating methoxy groups. In our case, it is noteworthy that, despite the absence of any activating groups and the choice of a lower metal loading (15 mol% vs 480 mol%), a higher conversion 44% ((120 °C, 2 h), MeCN - 10 vol% AcOH), vs 20% (90 °C, 20 min, acetone-water) and a higher selectivity for C-C cleavage (24% vs 18%) were obtained. However, Evtuguin et al. used acetone-water 70/30 v/v as a solvent which is less prone to solubilize O2 (1.2 mM [44,45]) compared to MeCN - 10 vol% AcOH (2.6 mM). Such result confirms that molecular oxygen transfer has a critical impact on aerobic cleavage and a reaction time of 20 min is too short to observe high yields of C-C cleavage.

4.4. Mechanism Proposal

In order to establish a plausible mechanism for K1HH cleavage, we also tried to identify the minor products by using GC-MS analysis. The different cleavage products detected are listed in Table S2 depending on their origin (C-O cleavage (Table S2a), C-C cleavage (Table S2b)). p-Quinone (from phenol oxidation) as well as phenylglyoxylic acid and 2-acetoxyacetophenone (from the acetolysis of K1HH, Table S2a) were identified. In addition, N-acetylbenzamide (Table S2b) could be detected. This amide would result from the reaction of benzoic acid with acetamide, the latter being produced by the hydrolysis of acetonitrile. Phenyl formate was also observed (Table S2b), meaning that formic acid maybe produced during the reaction.
The analysis of the undiluted sample allowed us to detect 1-(2-hydroxyphenyl)-2-phenoxyethanone and its isomer 2-(2-hydroxyphenyl)-1-phenoxyethanone (Table S2c). Both of them would be produced by the hydroxylation of the aromatic rings of K1HH by HO. radicals. Another compound (X) would be 2-phenoxy-2-acetoxy-1-phenylethanone (Figure 11 and Figure S10, Table S4).
A mechanistic pathway explaining the formation of the main products, other compounds being either intermediates or products from dead-end routes is proposed in Figure 12.
Hence, benzoic acid would be generated directly from K1HH through a direct C-C cleavage via a peroxyl radical intermediate (pathway (1)) [55], or through the C-O cleavage of a free radical intermediate Y3 (pathway (2)). Y3 could have reacted with acetic acid to form X (supposed from GC-MS data).
By analogy with cyclohexanone, studied by Cavani et al. [52], K1HH may be also tautomerized into its corresponding enol ether (Figure 12) as a result of an acid-catalyzed process involving H3+xPVx. One electron oxidation of the enol by vanadium (V) can be envisaged [13,56] affording X1 and X2 radicals. Then, X2 may react with molecular oxygen, giving rise to a hydroperoxyl radical, Y1 [49]. The latter would react with a hydrogen donor, for example, the substrate itself [52] affording a hydroperoxide Y2. It has to be noted that Y2 could also be produced by the reaction of O2.− with the X2 radical followed by a protonation step. Such superoxide anions would be formed in-situ by the one-electron oxidation of vanadium (+IV) by molecular oxygen [57]. Homolytic cleavage of the O-O bond in Y2 would occur afterwards, leading to Y3 then to phenylglyoxal and a phenoxy radical that can be reduced into phenol. Further oxidation of phenylglyoxal may give an α-keto carboxylic acid, phenylglyoxylic acid that can be decarboxylated, thus producing benzaldehyde and CO2. Phenylglyoxal may also be formed from 2-hydroxyacetophenone, the hydrolysis product of K1HH of which only traces were observed due to its high reactivity compared to K1HH as the result of the presence of the OH group [14,50].
Hydrolysis was shown indirectly by the presence of mandelic acid (PhCH(OH)COOH) obtained from the isomerization of 2-hydroxyacetophenone (PhCOCH2OH) into mandelaldehyde (PhCH(OH)CHO) and then aerobic oxidation [58]. Benzaldehyde cannot be formed through the pathway (1), that is, cleavage of the C-C bond into phenol and phenyl formate, which means that the pathway (2) and also the oxidation via the hydrolysis product (2-hydroxyacetophenone) take place. Yet, the formation of substantial amounts of phenyl formate shows that the oxidative cleavage occurs through pathway (1) too. Let us notice that i) the oxidation through pathway (1) and ii) an initial hydrolysis of K1HH and cleavage into PhCHO and CO2 consumes at least one equivalent of molecular oxygen whereas iii) the oxidation into PhOH and PhCHO through pathway (2) consumes at least 1.5 equivalents of molecular oxygen and one equivalent of carbon dioxide is emitted. So, the observed oxygen consumption (1.5 mol/converted K1HH) is adequate to this mechanism.
A similar search for the by-products was done also in the case of A1HH leading to a classification of the cleavage products arising either from C-O and C-C cleavage or products from side reactions (Table S2a–c). By analogy with K1HH, phenyl formate (C-C cleavage) and acetate (C-O cleavage), 2-acetoxyacetophenone (acetolysis) and phenyl benzoate were observed as minor products from A1HH cleavage but some other compounds characteristic of A1HH cleavage, such as phenylacetaldehyde and hydroxystilbenes were also evidenced. A first possible mechanism for A1HH cleavage would involve the oxidation of A1HH into K1HH, then the different pathways described in Figure 12. Indeed, K1HH was detected as a main product and 2-phenoxyacetophenone, as well as 2-phenoxy-2-acetoxy-1phenylethanone were also detected. However, the PhCHO/PhCOOH ratio in A1HH cleavage was shown to be higher than in K1HH cleavage which could be related to the involvement of another mechanism (Figure 13) inspired by Evtuguin et al. [27].
Two pathways were suggested—the first one (pathway (3)) involves a monoelectronic oxidation of the Ar1 moiety by the catalyst into a radial cation, Rad1. This activates the substrate for C-C cleavage into the Int1A cation and the radical, Rad1B. Then, Int1A would be deprotonated giving rise to benzaldehyde while Rad1B would be further oxidized (monoelectronic transfer) into the Int1B cation whose hydration leads to a formaldehyde hemiacetal. Finally, this one would be cleaved into phenol and formaldehyde. The second mechanism (pathway (4)) involves a monoelectronic oxidation of the Ar2 moiety into the radical cation, Rad2. Then, C-O cleavage in Rad2 would take place, affording the Rad2A radical and the Int2B cation. Int2B may be hydrated into hydroquinone while Rad2A is further oxidized (monoelectronic transfer) into the Int2A cation that is hydrated into styrenediol.
At first sight, styrene diol and hydroquinone were not detected by GC-MS, which would invalidate the pathway (4). So, the magnitude order of the PhCHO/PhCOOH ratio would be explained by A1HH oxidation according pathway (3) only. However, the formation of phenylacetaldehyde and acetophenone is not taken into account in Figure 12 and Figure 13. A priori, such products are generated by a non-oxidative C-O cleavage (Figure 14). Their common precursor would be the dehydration product of A1HH, that is, D1HH. Hence, D1HH would be rehydrated then into an unstable hemiacetal whose C-O cleavage gives rise to phenol and phenylacetaldehyde. The latter one may also isomerize into acetophenone in acidic media (Figure 14). Basically, such reaction would proceed through an acid-catalyzed H-shift [59], boosted by the enhanced stability of the R1 carbocation which in turn gives rise to a protonated epoxide R2 as a transient species. The ring is then opened, affording R3 and, after deprotonation and tautomerization, acetophenone. It has to be noticed that phenylacetaldehyde polymerizes easily, even at room temperature [60], that may be the reason why this aldehyde could not be detected by HPLC and GC-MS. As a consequence, it has to be emphasized that, on the contrary to esterification, the dehydration pathway cannot be considered at all as a dead-end route. However, it essentially gives rise to a non-oxidative C-O cleavage with a risk of polymerization.
Finally, hydroxystilbenes were shown to be minor side products of the oxidative condensation. Knowing that the C-O bond is the weakest linkage in lignin [61] and in its models too, we can suggest a homolytic cleavage (as depicted in Figure 15) affording one O- and one C- centered radicals. Then, inspired by lignification process [62] and knowing that repolymerized lignin contains a higher amount of C-C bonds compared to native lignin [61], the second step on the mechanism would be the isomerization of the phenoxy radical into a new C-centered radical and its homolytic coupling with the other one, previously formed. The hydroxystilbenes would be finally obtained through a dehydration step.

4.5. Extension to Lignin

It is well-known that lignin models have not really the reactivity of lignin. Therefore, the extension to lignin had to be addressed. The yields of C-C cleavage products from the Organosolv lignin were really low as shown in Table 4, even lower compared to A1HH cleavage. Similarly to the A1HH cleavage, the reason would be the esterification of the alcohol functions during the reaction. To get insight on the fate of the hydroxyl groups, a 31P NMR analysis of phosphorylated lignin (reaction given in Figure S11) samples before and after oxidation was performed. An example of spectrum of lignin before oxidation is given in Figure S12a. The signals related to lignin are given in Figure S12b and their attribution in Table S5. It can be noticed that the fractions of phenolic OH functions (21 vs 40%) decreased in the run performed without the H6PV3 catalyst (Table 4, entry 3) meaning that these functions were the most reactive, but for oxidative condensation since almost no C-C cleavage products was observed even after silylation. The drop of phenolic moieties content took place at a lesser extent in presence of catalyst meaning that a higher amount of aliphatic functions may be chemically transformed, which is compatible to a higher yield of C-C cleavage reactions as observed by GC-MS. In addition, we observed that oxidized lignin (120 °C, 2 h) either in the presence or without H6PV3 is much less soluble in dimethyl sulfoxide (DMSO) (2 g/L vs >20 g/L for lignin before oxidation) and in CDCl3-pyridine 1-1.6 v/v (after phosphorylation). The oxidized lignin (O2 5 bar, Mo + V 15 mol%, 120 °C, 2 h) was also tested for aerobic cleavage in the same operative conditions and showed to be resistant to cleavage and no dissolution was observed. This is due to the formation of more recalcitrant C-C bonds at the expense of C-O linkages [61] through condensation reactions as observed for A1HH cleavage but at a much higher extent because of the presence of phenolic groups [61,63].
It was noticed that 9 wt% of lignin was dissolved either with or without catalyst. The solubility cannot be due to the formation of monoaromatics only but also soluble oligomers formation from oxidative cleavage or acid-catalyzed C-O cleavage. The distribution of H (46%–50%), G (36%–38%) and S (12%–16%) monoaromatic compounds was rather unchanged, meaning that the different monomers had similar reactivities.

5. Conclusions

A Keggin-type molybdovanadophosphoric acid with the following mean formulae H5.7[PMo9.3V2.7O40] (H6PV3) was easily synthesized (79% yield) through a hydrothermal procedure and characterized by XRD, 31P NMR in D2O-H2O 50:50, ICP and TGA. The addition of 10 vol% of acetic acid to an oxidation resistant solvent, such as acetonitrile, was shown to be beneficial for the ketone cleavage mainly as the result of an enhanced molecular oxygen solubility. In the optimized conditions (82 °C, 24 h), the conversion of the ketone K1HH reached 72% and the yield of PhOH, so as the combined yield of PhCHO and PhCOOH were around 55%. Nevertheless, K1HH is quite far from lignin compared to the alcohol model A1HH. The latter one proved to be more resistant and the O2 pressure had to be increased. Besides, the acetate ester of A1HH, formed as a main by-product, turned out to be even more resistant than A1HH which indicates that esterification has to be mitigated. A tentative of optimization of the O2 pressure and of the reaction temperature was carried out for A1HH cleavage. Yet, at high pressure and temperature, the carbon balance tended to be low, implying that side reactions such as hydroxystilbenes formation through oxidative condensation and tar formation are involved. As a result, a temperature of 120 °C, a molecular oxygen pressure of 5 bar and a duration of 2 h were the most adequate parameters.
From the mechanistic studies, it was shown that, in K1HH, the active site is the CH2 group and that the C-C cleavage may proceed either through hydrolysis of K1HH or through the formation of a hydroperoxyl radical. The latter one may be cleaved directly into phenyl formate and benzoic acid or transformed into a radical from hydroxylated K1HH evoluting to phenol and phenylglyoxal that is further oxidized into phenylglyoxylic acid whose decarboxylation generates benzaldehyde. A1HH cleavage may be direct or proceed through alcohol oxidation into the ketone K1HH. Acidic C-O cleavage was evidenced as a side reaction.
Since the alcohol is more resistant to cleavage and the benzylic alcohol moiety (carbon Cα) is an active site for both dead-end oxidative condensation and esterification, a stepwise procedure may be a more suitable strategy [64] to cut down these side reactions. It was used not only for oxidative cleavage but also for reductive and acidic cleavage of lignin models [64,65,66,67]. For example, the first step would consist in oxidizing the benzylic alcohol into a ketone in presence of a catalytic system involving an oxidant such as TEMPO ((2,2,6,6-Tetramethylpiperidin-1-yl)oxidanyl) [68] and then, ketone cleavage should be performed in atmospheric conditions.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/13/4/812/s1, synthesis procedure of K1HH, A1HH and EstαA1HH, a general HPLC profile of the reaction media (Figure S1), the calibration curves of PhOH, PhCHO, PhCOOH, K1HH, A1HH and EstαA1HH (Figure S2), the list of products identified by HPLC (Table S1) and by GC-MS (Tables S2a–c), XRD profiles of H6PV3 vs starting MoO3 and V2O5 and reference H3PMo12, Figure S3) and a Rietveld refinement (Figure S4), graphical relaxation delay of H6PV3 (FigureS5 and S6), graphical calculation of the kinetic order to K1HH and A1HH (Figure S7), the calibration curves of phenolic aldehydes and acids from C-C cleavage of lignin (Figure S8), details on the calculation of oxygen solubility in binary acetonitrile and acetic acid (Figure S9), list of Organosolv lignin aerobic C-C cleavage identified by GC-MS (Table S3), identification of the structure of the intermediate X (Table S4 and Figure S10) and 31P NMR analysis of phosphorylated lignin (reaction detailed in Figure S11 and spectrum in Figure S12a and S12b, identification list on Table S5).

Author Contributions

Conceptualization, L.A.-H., F.L. and E.G.; methodology, L.A.-H. and F.L.; validation, F.L. and E.G.; formal analysis, L.A.-H.; investigation, L.A.-H., F.L. and E.G.; resources, F.L. and E.G.; data curation, L.A.-H., F.L. and E.G.; writing—original draft preparation, L.A.H.; writing—review and editing, L.A.-H., F.L. and E.G.; visualization, L.A.-H., F.L. and E.G.; supervision, F.L. and E.G.; project administration, L.A.-H., F.L. and E.G.; funding acquisition, F.L. and E.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by SORBONNE UNIVERSITE and CNRS (Program “Génie des Procédés”, Ph. D. fellowship of L.A.H. 2016-2019).

Acknowledgments

We acknowledge the Program “Génie des Procédés” for the Ph.D fellowship (2016-2019) of LAH. We would like to thank Mohamed SELMANE (XRD), Claire TROUFFLARD and Aurélie BERNARD (NMR), Martine TESSIER, Brigitte ROUSSEAU and Julien BRISSE (lignin characterization) for their technical support and our lignin suppliers Michel and Guo-Hua DELMAS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Behling, R.; Valange, S.; Chatel, G. Heterogeneous catalytic oxidation for lignin valorization into valuable chemicals: what results? What limitations? What trends? Green Chem. 2016, 18, 1839–1854. [Google Scholar] [CrossRef]
  2. Hocking, M.B. Vanillin: Synthetic Flavoring from Spent Sulfite Liquor. J. Chem. Educ. 1997, 74, 1055. [Google Scholar] [CrossRef]
  3. Lange, H.; Decina, S.; Crestini, C. Oxidative upgrade of lignin—Recent routes reviewed. Eur. Polym. J. 2013, 49, 1151–1173. [Google Scholar] [CrossRef] [Green Version]
  4. Ma, R.; Guo, M.; Zhang, X. Recent advances in oxidative valorization of lignin. Catal. Today 2018, 302, 50–60. [Google Scholar] [CrossRef]
  5. Mottweiller, J.; Puche, M.; Räuber, C.; Schmidt, T.; Concepción, P.; Corma, A.; Bolm, C. Copper- and Vanadium-Catalyzed Oxidative Cleavage of Lignin using Molecular oxygen. ChemSusChem 2015, 8, 2106–2113. [Google Scholar] [CrossRef]
  6. Díaz-Urrutia, C.; Hurisso, B.B.; Gauthier, P.M.; Sedai, B.; Singer, R.D.; Baker, R.T. Catalytic aerobic oxidation of lignin-derived bio-oils using oxovanadium and copper complex catalysts and ionic liquids. J. Mol. Catal. A Chem. 2016, 423, 414–422. [Google Scholar] [CrossRef]
  7. Amadio, E.; Di Lorenzo, R.; Zonta, C.; Licini, G.M. Vanadium catalyzed aerobic carbon–carbon cleavage. Co-ord. Chem. Rev. 2015, 301, 147–162. [Google Scholar] [CrossRef]
  8. Zhao, Y.; Xu, Q.; Pan, T.; Zuo, Y.; Fu, Y.; Guo, Q.-X. Depolymerization of lignin by catalytic oxidation with aqueous polyoxometalates. Appl. Catal. A Gen. 2013, 467, 504–508. [Google Scholar] [CrossRef]
  9. De Gregorio, G.F.; Prado, R.; Vriamont, C.; Erdocia, X.; Labidi, J.; Hallett, J.P.; Welton, T. Oxidative Depolymerization of Lignin Using a Novel Polyoxometalate-Protic Ionic Liquid System. ACS Sustain. Chem. Eng. 2016, 4, 6031–6036. [Google Scholar] [CrossRef]
  10. Cheng, F.; Wang, H.; Rogers, R.D. Oxygen Enhances Polyoxometalate-based Catalytic Dissolution and Delignification of Woody Biomass in Ionic Liquids. ACS Sustain. Chem. Eng. 2014, 2, 2859–2865. [Google Scholar] [CrossRef]
  11. Weinstock, I.A. Delignification of Wood Pulp by Vanadium-Substituted Polyoxometalates. US Patent 5302248, 12 April 1994. [Google Scholar]
  12. Gaspar, A.R.; Gamelas, J.A.F.; Evtuguin, D.V.; Neto, C.P. Alternatives for lignocellulosic pulp delignification using polyoxometalates and oxygen: a review. Green Chem. 2007, 9, 717. [Google Scholar] [CrossRef]
  13. Weinstock, I.A.; Schreiber, R.E.; Neumann, R. Molecular oxygen in polyoxometalate mediated reactions. Chem. Rev. 2018, 118, 2680–2717. [Google Scholar] [CrossRef] [PubMed]
  14. El Aakel, L.; Launay, F.; Atlamsani, A.; Brégeault, J.-M. Efficient and selective catalytic oxidative cleavage of α-hydroxy ketones using vanadium-based HPA and dioxygen. Chem. Commun. 2001, 21, 2218. [Google Scholar] [CrossRef] [PubMed]
  15. El Ali, B.; El-Ghanam, A.M.; Fettouhi, M. H3+nPMo12−nVnO40-catalyzed selective oxidation of benzoins to benzils or aldehydes and esters by molecular oxygen. J. Mol. Catal. A 2001, 165, 283–290. [Google Scholar] [CrossRef]
  16. Ballarini, N.; Casagrandi, L.; Cavani, F.; D’Alessandro, T.; Frattini, A.; Accorinti, P.; Alini, S.; Babini, P. The liquid-phase oxidation of cyclohexanone with oxygen, catalysed by Keggin-type polyoxometalates. A cleaner alternative to the current industrial process for adipic acid synthesis. DGMK Tagungsbericht 2008, 225–232. [Google Scholar]
  17. Rinaldi, R.; Jastrzebski, R.; Clough, M.T.; Ralph, J.; Kennema, M.; Bruijnincx, P.C.A.; Weckhuysen, B.M. Paving the Way for Lignin Valorisation: Recent Advances in Bioengineering, Biorefining and Catalysis. Angew. Chem. Int. Ed. 2016, 55, 8164–8215. [Google Scholar] [CrossRef] [Green Version]
  18. Sedai, B.; Díaz-Urrutia, C.; Baker, R.T.; Wu, R.; Silks, L.P.; Hanson, S.K. Comparison of Copper and Vanadium Homogeneous Catalysts for Aerobic Oxidation of Lignin Models. ACS Catal. 2011, 1, 794–804. [Google Scholar] [CrossRef]
  19. Hanson, S.K.; Baker, R.T.; Gordon, J.C.; Scott, B.L.; Thorn, D.L. Aerobic Oxidation of Lignin Models Using a Base Metal Vanadium Catalyst. Inorg. Chem. 2010, 49, 5611–5618. [Google Scholar] [CrossRef]
  20. Son, S.; Toste, F.D. Non-Oxidative Vanadium-Catalyzed C-O Bond Cleavage: Application to Degradation of Lignin Model Compounds. Angew. Chem. Int. Ed. 2010, 49, 3791–3794. [Google Scholar] [CrossRef] [Green Version]
  21. Zhang, G.; Scott, B.L.; Wu, R.; Silks, L.P.; Hanson, S.K. Aerobic Oxidation Reactions Catalyzed by Vanadium Complexes of Bis(Phenolate) Ligands. Inorg. Chem. 2012, 51, 7354–7361. [Google Scholar] [CrossRef]
  22. Hanson, S.K.; Wu, R.; Silks, L.P. C-C or C-O Bond Cleavage in a Phenolic Lignin Model Compound: Selectivity Depends on Vanadium Catalyst. Angew. Chem. Int. Ed. 2012, 51, 3410–3413. [Google Scholar] [CrossRef] [PubMed]
  23. Ma, Y.; Du, Z.; Liu, J.; Xia, F.; Xu, J. Selective oxidative C–C bond cleavage of a lignin model compound in the presence of acetic acid with a vanadium catalyst. Green Chem. 2015, 17, 4968–4973. [Google Scholar] [CrossRef]
  24. Jiang, Y.-Y.; Yan, L.; Yu, H.-Z.; Zhang, Q.; Fu, Y. Mechanism of Vanadium-Catalyzed Selective C–O and C–C Cleavage of Lignin Model Compound. ACS Catal. 2016, 6, 4399–4410. [Google Scholar] [CrossRef]
  25. Parker, H.J.; Chuck, C.J.; Woodman, T.; Jones, M.D. Degradation of β-O-4 model lignin species by vanadium Schiff-base catalysts: Influence of catalyst structure and reaction conditions on activity and selectivity. Catal. Today 2016, 269, 40–47. [Google Scholar] [CrossRef] [Green Version]
  26. Gazi, S.; Đokić, M.; Moeljadi, A.M.P.; Ganguly, R.; Hirao, H.; Soo, H.S. Kinetics and DFT Studies of Photoredox Carbon–Carbon Bond Cleavage Reactions by Molecular Vanadium Catalysts under Ambient Conditions. ACS Catal. 2017, 7, 4682–4691. [Google Scholar] [CrossRef]
  27. Evtuguin, D.V.; Daniel, A.I.; Silvestre, A.J.; Amado, F.M.; Neto, C.P. Lignin aerobic oxidation promoted by molybdovanadophosphate polyanion [PMo7V5O40]8−. Study on the oxidative cleavage of β-O-4 aryl ether structures using model compounds. J. Mol. Catal. A Chem. 2000, 154, 217–224. [Google Scholar] [CrossRef]
  28. Kozhevnikov, I.V. Catalysis by heteropoly acids and multicomponent polyoxometalates in liquid-phase reactions. Chem. Rev. 1998, 98, 171–198. [Google Scholar] [CrossRef]
  29. Bujanovic, B.; Ralph, S.; Reiner, R.; Hirth, K.; Atalla, R. Polyoxometalates in Oxidative Delignification of Chemical Pulps: Effect on Lignin. Materials 2010, 3, 1888–1903. [Google Scholar] [CrossRef] [Green Version]
  30. Odyakov, V.F.; Zhizhina, E.G.; Rodikova, Y.A.; Gogin, L.L. Mo-V-Phosphoric Heteropoly Acids and Their Salts: Aqueous Solution Preparation—Challenges and Perspectives. Eur. J. Inorg. Chem. 2015, 2015, 3618–3631. [Google Scholar] [CrossRef]
  31. Kern, F.; Ruf, S.; Emig, G. Vapour-phase trimerization of formaldehyde to trioxane catalysed by 1-vanado-11-molybdophosphoric acid. Appl. Catal. A Gen. 1997, 150, 143–151. [Google Scholar] [CrossRef]
  32. Tsigdinos, G.A.; Hallada, C.J. Molybdovanadophosphoric acids and their salts. I. Investigation of methods of preparation and characterization. Inorg. Chem. 1968, 7, 437–441. [Google Scholar] [CrossRef]
  33. Derwent, R.; Jenkin, M.; Passant, N.; Pilling, M. Reactivity-based strategies for photochemical ozone control in Europe. Environ. Sci. Policy 2007, 10, 445–453. [Google Scholar] [CrossRef]
  34. Atlamsani, A.; Ziyad, M.; Brégeault, J. Les hétéropolyacides comme catalyseurs bifonctionnels pour la coupure oxydante de cyclanones. J. Chim. Phys. 1995, 92, 1344–1364. [Google Scholar] [CrossRef]
  35. Odyakov, V.F.; Zhizhina, E.G. New process for preparing aqueous solutions of Mo-V-phosphoric heteropoly acids. Russ. J. Inorg. Chem. 2009, 54, 361–367. [Google Scholar] [CrossRef]
  36. Grate, J.H.; Hamm, D.R.; Mahajan, S. Palladium and phosphomolybdovanadate catalyzed olefin oxidation to carbonyls. Mol. Eng. 1993, 3, 205–229. [Google Scholar] [CrossRef]
  37. Delmas, M.; Benjelloun-Mlayah, B. Procédé de Prétraitement d’une Matière Végétale Lignocellulosique en vue de la Production de Bioethanol; Pat. Fr 2 926 824; INPI: Paris, France, 21 January 2008. [Google Scholar]
  38. Mbotchak, L.; Le Morvan, C.; Duong, K.L.; Rousseau, B.; Tessier, M.; Fradet, A. Purification, Structural Characterization, and Modification of Organosolv Wheat Straw Lignin. J. Agric. Food Chem. 2015, 63, 5178–5188. [Google Scholar] [CrossRef]
  39. Marchal-Roch, C.; Bayer, R.; Moisan, J.F.; Teze, A.; Hervé, G. Oxidative dehydrogenation of isobutyric acid: Characterization and modeling of vanadium containing polyoxometalate catalysts. Top. Catal. 1996, 3, 407–419. [Google Scholar] [CrossRef]
  40. Pettersson, L.; Andersson, I.; Selling, A.; Grate, J.H. Multicomponent polyanions. 46. Characterization of the isomeric Keggin decamolybdodivanadophosphate ions in aqueous solution by 31P and 51V NMR. Inorg. Chem. 1994, 33, 982–993. [Google Scholar] [CrossRef]
  41. Barats-Damatov, D.; Shimon, L.J.W.; Feldman, Y.; Bendikov, T.; Neumann, R. Solid-State Crystal-to-Crystal Phase Transitions and Reversible Structure–Temperature Behavior of Phosphovanadomolybdic Acid, H5PV2Mo10O40. Inorg. Chem. 2015, 54, 628–634. [Google Scholar] [CrossRef] [PubMed]
  42. Villabrille, P.; Romanelli, G.; Gassa, L.; Vazquez, P.; Caceres, 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]
  43. Zhao, X.; Zhang, Y.; Wei, L.; Hu, H.; Huang, Z.; Yang, M.; Huang, A.; Wu, J.; Feng, Z. Esterification mechanism of lignin with different catalysts based on lignin model compounds by mechanical activation-assisted solid-phase synthesis. RSC Adv. 2017, 7, 52382–52390. [Google Scholar] [CrossRef] [Green Version]
  44. Sato, T.; Hamada, Y.; Sumikawa, M.; Araki, S.; Yamamoto, H. Solubility of Oxygen in Organic Solvents and Calculation of the Hansen Solubility Parameters of Oxygen. Ind. Eng. Chem. Res. 2014, 53, 19331–19337. [Google Scholar] [CrossRef]
  45. Quaranta, M.; Murkovic, M.; Klimant, I. A new method to measure oxygen solubility in organic solvents through optical oxygen sensing. Analyst 2013, 138, 6243. [Google Scholar] [CrossRef] [PubMed]
  46. Wu, X.; Deng, Z.; Yan, J.; Zhang, Z.; Zhang, F.; Zhang, Z. Experimental Investigation on the Solubility of Oxygen in Toluene and Acetic Acid. Ind. Eng. Chem. Res. 2014, 53, 9932–9937. [Google Scholar] [CrossRef]
  47. Khenkin, A.M.; Neumann, R. Oxidative C−C Bond Cleavage of Primary Alcohols and Vicinal Diols Catalyzed by H5PV2Mo10O40 by an Electron Transfer and Oxygen Transfer Reaction Mechanism. J. Am. Chem. Soc. 2008, 130, 14474–14476. [Google Scholar] [CrossRef]
  48. Stadler, R.H.; Welti, D.H.; Stämpfli, A.A.; Fay, L.B. Thermal Decomposition of Caffeic Acid in Model Systems: Identification of Novel Tetraoxygenated Phenylindan Isomers and Their Stability in Aqueous Solution. J. Agric. Food Chem. 1996, 44, 898–905. [Google Scholar] [CrossRef]
  49. Neumann, R.; Khenkin, A.M. Molecular oxygen and oxidation catalysis by phosphovanadomolybdates. Chem. Commun. 2006, 24, 2529. [Google Scholar] [CrossRef]
  50. Brégeault, J.-M. Transition-metal complexes for liquid-phase catalytic oxidation: some aspects of industrial reactions and of emerging technologies. Dalton Trans. 2003, 17, 3289–3302. [Google Scholar] [CrossRef]
  51. El Amrani, I.; Atlamsani, A.; Dakkach, M.; Rodríguez, M.; Romero, I.; Amthiou, S. Efficient and selective oxidation of aldehydes with dioxygen catalysed by vanadium-containing heteropolyanions. Comptes Rendus Chim. 2017, 20, 888–895. [Google Scholar] [CrossRef]
  52. Cavani, F.; Ferroni, L.; Frattini, A.; Lucarelli, C.; Mazzini, A.; Raabova, K.; Alini, S.; Accorinti, P.; Babini, P. Evidence for the presence of alternative mechanisms in the oxidation of cyclohexanone to adipic acid with oxygen, catalysed by Keggin polyoxometalates. Appl. Catal. A Gen. 2011, 391, 118–124. [Google Scholar] [CrossRef]
  53. Muckerman, J.T.; Skone, J.H.; Ning, M.; Wasada-Tsutsui, Y. Toward the accurate calculation of pKa values in water and acetonitrile. Biochim. Biophys. Acta (BBA) Gen. Subj. 2013, 1827, 882–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Fujioka, G.S.; Cady, G.H. Inorganic Chemistry in Trifluoroacetic Acid1. J. Am. Chem. Soc. 1957, 79, 2451–2454. [Google Scholar] [CrossRef]
  55. Bailey, P.S.; Chang, Y.-G. Rearrangements of Certain α-Alkoxy-α-hydroperoxyacophenones. Alcoholysis of Acetic Benzoic Anhydride. J. Org. Chem. 1962, 27, 1192–1197. [Google Scholar] [CrossRef]
  56. Efremenko, I.; Neumann, R. Computational Insight into the Initial Steps of the Mars–van Krevelen Mechanism: Electron Transfer and Surface Defects in the Reduction of Polyoxometalates. J. Am. Chem. Soc. 2012, 134, 20669–20680. [Google Scholar] [CrossRef] [PubMed]
  57. Geletii, Y.V.; Hill, C.L.; Atalla, R.H.; Weinstock, I.A. Reduction of O2 to Superoxide Anion (O2•−) in Water by Heteropolytungstate Cluster-Anions. J. Am. Chem. Soc. 2006, 128, 17033–17042. [Google Scholar] [CrossRef] [PubMed]
  58. De Gregorio, G.F.; Weber, C.C.; Gräsvik, J.; Welton, T.; Brandt, A.; Hallett, J.P. Mechanistic insights into lignin depolymerisation in acidic ionic liquids. Green Chem. 2016, 18, 5456–5465. [Google Scholar] [CrossRef]
  59. Tian, Y.; Jürgens, E.; Mill, K.; Jordan, R.; Maulbetsch, T.; Kunz, D. Nucleophilic Isomerization of Epoxides by Pincer-Rhodium Catalysts: Activity Increase and Mechanistic Insights. ChemCatChem 2019, 11, 4028–4035. [Google Scholar] [CrossRef] [Green Version]
  60. Kuhn, W. Method for stabilizing phenylacetaldehyde. US Patent 6 624 330, 23 September 2003. [Google Scholar]
  61. Schutyser, W.; Renders, T.; Bosch, S.V.D.; Koelewijn, S.-F.; Beckham, G.T.; Sels, B.F. Chemicals from lignin: an interplay of lignocellulose fractionation, depolymerisation, and upgrading. Chem. Soc. Rev. 2018, 47, 852–908. [Google Scholar] [CrossRef]
  62. Freudenberg, K. Biosynthesis and constitution of lignin. Nature 1959, 183, 1152–1155. [Google Scholar] [CrossRef]
  63. Akim, L.G.; Colodette, J.L.; Argyropoulos, D.S. Factors limiting oxygen delignification of kraft pulp. Can. J. Chem. 2001, 79, 201–210. [Google Scholar] [CrossRef]
  64. Yu, X.; Wei, Z.; Lu, Z.; Pei, H.; Wang, H. Activation of lignin by selective oxidation: An emerging strategy for boosting lignin depolymerization to aromatics. Bioresour. Technol. 2019, 291, 121885. [Google Scholar] [CrossRef] [PubMed]
  65. Magallanes, G.; Kärkäs, M.D.; Bosque, I.; Lee, S.; Maldonado, S.; Stephenson, C.R.J. Selective C–O Bond Cleavage of Lignin Systems and Polymers Enabled by Sequential Palladium-Catalyzed Aerobic Oxidation and Visible-Light Photoredox Catalysis. ACS Catal. 2019, 9, 2252–2260. [Google Scholar] [CrossRef]
  66. Liu, C.; Wu, S.; Zhang, H.; Xiao, R. Catalytic oxidation of lignin to valuable biomass-based platform chemicals: A review. Fuel Process. Technol. 2019, 191, 181–201. [Google Scholar] [CrossRef]
  67. Wang, Y.; Wang, Q.; He, J.; Zhang, Y. Highly effective C–C bond cleavage of lignin model compounds. Green Chem. 2017, 19, 3135–3141. [Google Scholar] [CrossRef]
  68. Wang, M.; Lu, J.; Zhang, X.; Li, L.; Li, H.; Luo, N.; Wang, F. Two-Step, Catalytic C–C Bond Oxidative Cleavage Process Converts Lignin Models and Extracts to Aromatic Acids. ACS Catal. 2016, 6, 6086–6090. [Google Scholar] [CrossRef]
Figure 1. β-O-4′ bond in lignin polymer.
Figure 1. β-O-4′ bond in lignin polymer.
Materials 13 00812 g001
Figure 2. X-ray diffraction (XRD) profile of H6PV3 (main characteristic peaks marked by * symbol).
Figure 2. X-ray diffraction (XRD) profile of H6PV3 (main characteristic peaks marked by * symbol).
Materials 13 00812 g002
Figure 3. 31P nuclear magnetic resonance (NMR) profiles of V3.
Figure 3. 31P nuclear magnetic resonance (NMR) profiles of V3.
Materials 13 00812 g003
Figure 4. Thermogravimetric analysis (TGA)–differential scanning calorimetry (DSC) profile of H6PV3.
Figure 4. Thermogravimetric analysis (TGA)–differential scanning calorimetry (DSC) profile of H6PV3.
Materials 13 00812 g004
Figure 5. Cleavage of K1HH by O2.
Figure 5. Cleavage of K1HH by O2.
Materials 13 00812 g005
Figure 6. Evolution of different reaction parameters vs time for the H6PV3-catalyzed aerobic cleavage of K1HH; (a) yield of cleavage products, (b) carbon balance and selectivity for C-O and C-C cleavage products.
Figure 6. Evolution of different reaction parameters vs time for the H6PV3-catalyzed aerobic cleavage of K1HH; (a) yield of cleavage products, (b) carbon balance and selectivity for C-O and C-C cleavage products.
Materials 13 00812 g006aMaterials 13 00812 g006b
Figure 7. Aerobic oxidation of A1HH by O2.
Figure 7. Aerobic oxidation of A1HH by O2.
Materials 13 00812 g007
Figure 8. Evolution of different reaction parameters vs time for the H6PV3-catalyzed aerobic cleavage of A1HH; (a) yield of cleavage products, (b) carbon balance and selectivity for C-O and C-C cleavage products (see conditions in Table 3).
Figure 8. Evolution of different reaction parameters vs time for the H6PV3-catalyzed aerobic cleavage of A1HH; (a) yield of cleavage products, (b) carbon balance and selectivity for C-O and C-C cleavage products (see conditions in Table 3).
Materials 13 00812 g008
Figure 9. Aerobic oxidative cleavage of lignin (WSLp) catalyzed by H6PV3.
Figure 9. Aerobic oxidative cleavage of lignin (WSLp) catalyzed by H6PV3.
Materials 13 00812 g009
Figure 10. 31P NMR spectra of the fresh and spent H6PV3 catalysts (catalyst 30 mg in D2O-H2O 50:50 500 µL + dioxane 7.5 µL).
Figure 10. 31P NMR spectra of the fresh and spent H6PV3 catalysts (catalyst 30 mg in D2O-H2O 50:50 500 µL + dioxane 7.5 µL).
Materials 13 00812 g010
Figure 11. Hypothetic structure of the intermediate X.
Figure 11. Hypothetic structure of the intermediate X.
Materials 13 00812 g011
Figure 12. Mechanism proposal for the H6PV3 catalyzed aerobic cleavage of K1HH.
Figure 12. Mechanism proposal for the H6PV3 catalyzed aerobic cleavage of K1HH.
Materials 13 00812 g012
Figure 13. Mechanisms suggested for the direct cleavage of A1HH inspired by EVTUGUIN et al.
Figure 13. Mechanisms suggested for the direct cleavage of A1HH inspired by EVTUGUIN et al.
Materials 13 00812 g013
Figure 14. Suggested mechanisms for the formation of phenylacetaldehyde and acetophenone.
Figure 14. Suggested mechanisms for the formation of phenylacetaldehyde and acetophenone.
Materials 13 00812 g014
Figure 15. Suggested mechanisms for the formation of hydroxystilbenes.
Figure 15. Suggested mechanisms for the formation of hydroxystilbenes.
Materials 13 00812 g015
Table 1. Cell parameters of H6PV3 catalysts (vs reference H3PMo12) calculated on Fullprof® by the RIETVELD method.
Table 1. Cell parameters of H6PV3 catalysts (vs reference H3PMo12) calculated on Fullprof® by the RIETVELD method.
Materiala (Å)b (Å)c (Å)α (°)β (°)γ (°)RP (%)Chi2 (%)
V314.21914.41013.616112.54110.1360.1315.113.1
H3PMo12 (JCPDS 00-043-0317)14.10014.13013.550112.10109.8060.70--
Table 2. Influence of acetic acid content on the activity of H6PV3 on the K1HH aerobic oxidative cleavage.
Table 2. Influence of acetic acid content on the activity of H6PV3 on the K1HH aerobic oxidative cleavage.
Entryz
(vol%)
Mo + V
Loading
(mol%)
Conv.
(%)
Yield (%)O2 sol. (%) aCB
(%)
O2 Consumption b
PhOHPhCHOPhCOOH
1c015148.1042.1 d91n. i.
201548322.5192.3 d760.9
310015100800807.8 d740
42.515593417282.36782.0
5515684720382.42812.9e
61015725512442.56801.5
7f101544245.6242.56811.5
8201580670472.85721.4
9g1003935110.92.56821.3
1010743368242.56881.4
11103688472.2502.56581.1
K1HH 100 mM, atm. O2, H6PV3 (Mo + V 15 mol%), MeCN − z vol% AcOH, 82 °C, 24 h; n. d. = not detected. a The solubility in mixed solvents was estimated using equations from Ref 44; b Normalized to conversion; c Methanol (65 °C); d Obtained from Refs. [45,46]; e Due to experimental error; f T = 65 °C; g H2SO4 (5 µL) was used instead of V3.
Table 3. Activity of H6PV3 catalyst for A1HH cleavage.
Table 3. Activity of H6PV3 catalyst for A1HH cleavage.
EntryT
(°C)
P
(bar)
t
(h)
Conv.
(%)
Yield (%)CB
(%)
PhOHPhCHOPhCOOHK1HHEstαA1HHD1HH
180124354.62.02.117913108
2a801249.11.92.0trtr05.298
38052454167.19.8101821118
48056333.72.93.58.79.91099
512052447.16.47.52.7111696
612056907.62.6196.7121255
A1HH 100 mM, H6PV3 (Mo + V 15 mol%), MeCN - 10 vol% AcOH; a Without catalyst (in presence of H2SO4 6.2 mol%).
Table 4. H6PV3 catalyzed lignin aerobic cleavage.
Table 4. H6PV3 catalyzed lignin aerobic cleavage.
Entryt
(h)
Liquid Phase
(GC-MS after Silylation)
Residual Lignin (31P NMR)
Yield of Monoaromatics (mol%) bFunction RatioMonoaromatic Type Proportions b
Units HUnits GUnits SAliph./ArOH/COOHHGS
HHAVVASSA
120.40.11.20.30.2040/30/30163648
2 a2000.100.2039/21/40123850
360.60.30.70.40.30n. d.n. d.
WSLp-------36/40/24163846
WSLp 0.85 g (equivalent to 5 mmol of monoaromatic compounds); H6PV3 (Mo + V 15.9 mol% vs the estimated amount of dimeric units), 120 °C, O2 5 bar; a Without any catalyst, b Calculated from 31P NMR data.

Share and Cite

MDPI and ACS Style

Al-Hussaini, L.; Launay, F.; Galvez, E. Vanadium-Substituted Phosphomolybdic Acids for the Aerobic Cleavage of Lignin Models—Mechanistic Aspect and Extension to Lignin. Materials 2020, 13, 812. https://doi.org/10.3390/ma13040812

AMA Style

Al-Hussaini L, Launay F, Galvez E. Vanadium-Substituted Phosphomolybdic Acids for the Aerobic Cleavage of Lignin Models—Mechanistic Aspect and Extension to Lignin. Materials. 2020; 13(4):812. https://doi.org/10.3390/ma13040812

Chicago/Turabian Style

Al-Hussaini, Louay, Franck Launay, and Elena Galvez. 2020. "Vanadium-Substituted Phosphomolybdic Acids for the Aerobic Cleavage of Lignin Models—Mechanistic Aspect and Extension to Lignin" Materials 13, no. 4: 812. https://doi.org/10.3390/ma13040812

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