Wei-Xuan
Shu
a,
Lin-Yan
Bao
a,
Li-Li
Wang
a,
Xiao-Xia
You
a,
Xin
Ma
a,
Ya
Wang
*bc,
Zhong-Min
Su
*a and
Rong-Lin
Zhong
*a
aState Key Laboratory of Supramolecular Structure and Materials, Institute of Theoretical Chemistry, College of Chemistry, Jilin University, Changchun 130021, China. E-mail: suzhongmin@jlu.edu.cn; zhongrl898@jlu.edu.cn
bInstitute of Chemical & Industrial Bio-engineering, Jilin Engineering Normal University, Changchun, Jilin 130052, China. E-mail: wangy100@jlenu.edu.cn
cJilin Science and Technology Innovation Center of Green Synthesis and New Materials Research and Development, Changchun, Jilin 130052, China. E-mail: wangy100@jlenu.edu.cn
First published on 3rd June 2025
Polyoxovanadates (POVs) show great promise in catalytic oxidative desulphurisation, an industrially important yet challenging process, in which peroxovanadate generated by the oxidation of POVs is an essential active species. In such reaction, H2O2 is generally used as an oxidant because it is environmentally friendly. However, the formation mechanism of peroxovanadate is complicated, and the origin of its high reactive activity in sulfide oxidation remains unclear. Herein, we carefully investigate the generation mechanism of peroxovanadate active species formed from the reaction between POVs and H2O2 by density functional theory (DFT) calculations. The results clearly show that POV is oxidized by H2O2 to afford the highly reactive oxidant peroxovanadate intermediate via a proton transfer pathway rather than a radical pathway. Interestingly, the oxidation of VOSO4 by H2O2 affording a peroxovanadate intermediate differs from that of POVs, which likely occurs via a HO–OH bond cleavage (radical) pathway. On the other hand, the high activity of peroxovanadates in sulfide oxidation originates from the lower-energy empty orbitals of the oxygen atoms compared to those in POVs. These findings may provide a new perspective for the application of POV-based materials in the catalytic field.
Polyoxometalates (POMs) are widely recognized as environmentally friendly catalysts.15–20 POMs are a large class of metal (e.g., MoVI, WVI, VV, NbV, and TaV) oxides and represent a tremendous range of crystalline inorganic clusters with unparalleled physical and chemical properties.21 It is also noteworthy that the properties of POMs, including solubility, redox ability, acidity, magnetism, and thermal/chemical stability, can be tuned by modifying the constituent atoms, organic bridging components, and structural dimensions (such as POM-based chains, layers, and frameworks).22–29 Previously, some classic POMs have been extensively investigated as homogeneous catalysts in the selective oxidation of various sulfides.30–32 On the surface of POMs, numerous oxygen atoms act as proton receptors.33–36 Therefore, POMs and their derivatives can facilitate proton transfer to promote catalytic activation. Polyoxovanadates (POVs), a significant branch of POMs have variable coordination geometries of V atoms ({VO4} tetrahedral, {VO5} tetragonal cone, and {VO6} octahedral), which contribute to their structural diversity. Furthermore, the versatile oxidation states of vanadium (III, IV, and V) contribute to their unique redox properties, making them promising candidates for catalyzing oxidation reactions of small organic molecules.37 It is worthy of note that POVs have attracted considerable attention due to their efficient catalytic activity and potential applications in biomedicine and nano-magnetism.38,39
POVs are widely employed as catalysts or catalyst precursors for various oxidative reactions involving H2O2, tBuOOH (TBHP) or O2 under mild conditions.40 Recently, several experimental studies have reported on the catalytic activity of POVs in oxidative desulphurization reactions.41,42 Xu's group synthesized four methoxy and organic amine-functionalized windmill-shaped {V8} clusters for the oxidation of diphenyl thioether, resulting in diphenyl sulfoxide.15 Niu and coworkers reported a high-nuclearity POV-based aliphatic carboxylate derivative, K6H[V17VV12IV(OH)4O60(OOC(CH2)4COO)8]·nH2O,41 which also demonstrated catalytic activity for the oxidation of diphenyl thioether using TBHP as an oxidant. Recently, our group synthesized a series of POV-based metal–organic polyhedra with highly efficient catalytic activity for the oxidation of sulphur-containing substrates to sulfoxides and over-oxidized sulfones,43–48 showing good performance under milder reaction conditions. For instance, the reaction conditions of the {V5O9Cl}-based metal–organic polyhedral catalytic system are moderate, with methanol (CH3OH) as the solvent. Most notably, tert-butyl hydroperoxide (TBHP), an environment-friendly oxidant, was used instead of more aggressive chemicals, as shown in Fig. 1(a).48 The metal–organic polyhedral [(V5O9Cl)3(TDA)6]6− consists of three {V5O9Cl} units connected by a six thiophene dicarboxylic bridge (TDA), where each {V5O9Cl} unit contains four VIV atoms and one VV atom, five terminal oxygen atoms (Ot) and four bridge oxygen atoms (Ob) (Fig. 1(b)). According to previous investigations,49–51 in this system, the oxidation of a POV unit by TBHP likely generates a V-peroxo intermediate as active species, by which a sulfur containing substrate is oxidized to sulfoxides or sulfones.
The experimental results provide an environmentally friendly strategy for oxidative desulphurization. However, further theoretical studies are needed to better understand the formation mechanism of POV and the origin of its high activity in sulfide oxidation. Specifically, three questions remain regarding the experimental findings: (i) during the oxidation of {V5O9Cl}, both H–OOH and HO–OH bond cleavage of H2O2 might occur (Scheme 1). Which is the more favorable pathway for the oxidation process catalyzed by POV clusters? (ii) What role does the solvent (methanol) molecule play in the catalytic reaction? Vanadium oxovanadium sulphate/hydrochloride (VOSO4·nH2O/VOCl2·nH2O) exhibits some catalytic activity, but its catalytic performance is significantly enhanced when converted to the cluster form, as shown in Fig. 1(c). (iii) What is the key factor highlighting the advantages of the cluster? Therefore, it is crucial to disclose the origin of the high catalytic performance of the POV-based cluster. Answering these questions will provide the fundamental understanding and valuable information for the development of more excellent catalysts.
In this work, the mechanism of desulfurization oxidation catalyzed by [(V5O9Cl)3(TDA)6]6− and the key factors dominating the catalytic performance of {V5O9Cl} cluster were theoretically investigated by using density functional theory (DFT) calculations. The entire catalytic cycle consists of three elementary steps: activation of H2O2, generation of peroxovanadate, and oxygen atom transfer (OAT) from peroxovanadate to thioether. In this case, the preliminary two steps occur via the proton transfer pathway, facilitated by a hydrogen bonding network formed between the solvent molecule (methanol), H2O2 and POV cluster, which decreases the energy barrier. Differences in reactive behavior of VOSO4 and POV with H2O2 are mainly due to their different coordination environments of vanadium atoms, which lead to different pathways of H2O2 activation (via O–O or H–O bond cleavage). Furthermore, the study reveals that the molecular cage effect, formed by the six TDAs, results in a slight decrease in the energy barrier. The results clearly disclose that the high reactivity of peroxovanadate for sulfide oxidation originates from the lower energy of its empty orbitals compared to that in polyoxovanadates. These findings provide new perspectives on the application of POV-based catalytic materials in improving the environment.
Under experimental conditions, [(V5O9Cl)3(TDA)6] was employed as the catalyst for the H2O2 oxidative desulfurization reaction,59 with the three V5O9Cl clusters serving as the active units. Therefore, we explore the reaction by a representative cluster model V5O9Cl(TDA)4 to reduce the complexity and reduce the computational cost, as shown in Fig. 1(b). According to the crystal structure of [(V5O9Cl)3(TDA)6]6−, it can be inferred that one VV and four VIV atoms exist in the model cluster [V5O9Cl(TDA)4]2−. Furthermore, as shown in Fig. S4–S7,† the spin and SOMOs indicate that the oxidative state of four peripheral V atoms is VIV and the center one is in VV.
![]() | ||
Fig. 2 Gibbs energy profile (in kcal mol−1) of VOSO4 catalyzed methyl phenyl sulfoxide oxidation. The red dashed line represents the transition state part. |
However, VOSO4 may encounter the challenge of competitive decomposition of H2O2, as discussed in detail in Fig. S10.† Computational results suggest that VOSO4 shows catalytic activity for H2O2 decomposition, hence only a little portion of H2O2 is involved in the oxidation process of sulfide substrate, resulting in the lower conversion rate of sulfides.
Herein, we aim to briefly discuss the result that the activation energy of H2O2 activation by VOSO4 occurs via HO–OH bond cleavage, with this value being significantly lower than that for H–OOH bond cleavage. In VOSO4, the V atom has an unsaturated coordination environment and H2O2 interacts with VOSO4 to form intermediate II, exhibiting a considerable binding energy (−7.3 kcal mol−1). The oxidation of VVI to VV thus occurs via a concerted transition state TSIA, exothermically (−12.7 kcal mol−1) generating intermediate III with two hydroxyl groups bonding to the VV atom, a step with an energy barrier of 11.2 kcal mol−1. In addition, the path (TSIB and TSIC) via the H–OOH bond cleavage has higher energy barriers of 26.5 kcal mol−1 and 14.6 kcal mol−1. Thus, VOSO4 activation of H2O2 is via HO–OH bond cleavage. As illustrated in Fig. S11,† the frontier molecular orbital analysis shows that the interaction between the d orbital of V and the σ* orbital of the O–O bond occurs in a way that the p orbital of one OH group forms a bonding interaction with V, while the other one does not. Nevertheless, the inherent instability of the free hydroxyl group and unsaturated coordination of the V atom lead to the formation of intermediate IIIA, which converts to peroxovanadium sulphate though the dissociation of the H2O molecule.
According to the above theoretical studies on VOSO4 catalyzed oxidation of sulfides, the formation of peroxovanadium is important, and two plausible reaction mechanisms seem to exist for the oxidation process where POV is converted into the peroxovanadate active species, as shown in Scheme 2. In the first mechanism, H2O2 is activated via an O–O bond cleavage, forming a hydroxyl-containing V intermediate (V–OH) and a ˙OH free radical. This intermediate then undergoes hydrogen atom transfer dehydration, converting POV-based V5 clusters into the peroxovanadate intermediate. In the other mechanism, H2O2 is activated via a H–OOH bond cleavage, generating a peroxy V intermediate (V–OOH). The peroxovanadate intermediate is subsequently generated via proton transfer, which leads to the cleavage of the O–O bond and dehydration from the V–OOH intermediate. Next, the sulfide is oxidized by the peroxovanadate intermediate to form sulfoxide, with the oxidation proceeding through an oxygen atom transfer pathway.
![]() | ||
Scheme 2 Reaction mechanism of POV-catalyzed methyl phenyl sulfoxide oxidation. The organic ligands of the catalyst are omitted for clarity. |
The Gibbs energy profile of H2O2 activated though the O–O bond cleavage is shown in Fig. 3. H2O2 approaches [V5O9Cl(TDA)4]2− (1) to form the adduct [(V5O9Cl)(TDA)4⋯H2O2]2– (2A) with a considerable binding energy (−4.7 kcal mol−1) due to the formation of two hydrogen bonds (Ob⋯H and Ot1⋯H) with a distance of 1.90 Å. Subsequently, the transformation from 2A to 2B is endothermic (9.4 kcal mol−1) because it involves a deformation of the V5O9 cluster to ensure the reaction with H2O2. Specifically, the two V–Ob bonds of V5O9 elongate from 1.85 Å to 2.02 Å. It is important to note that the energy barriers will be particularly high (49.7 kcal mol−1), if the V5O9 cluster does not undergo deformation and is directly activated viaTS1A. Next, H2O2 coordinates to the V atom in 2B because the V–Oh1 distance is 4.30 Å and the Oh1–Oh2 distance is 1.46 Å, which are almost the same as those in the equilibrium structure. From 2B, the O–O activation of H2O2 proceeds viaTS1B, forming hydroxyl-containing intermediate 3B and a ˙OH free radical. The V–Oh1 decreases to 1.95 Å in TS1B, and Oh1–Oh2 elongates to 1.77 Å in TS1B. The V–Oh1 distance further decreases to 1.75 Å in 3B accompanied by the elongation of the Oh1–Oh2 bond (4.47 Å). The value of ΔG°‡ in this elementary step is 35.1 kcal mol−1 with a somewhat endothermic reaction energy (ΔG° = 2.2 kcal mol−1). After the dissociation of V–Ob, a similar pathway to the activation of H2O2 by VOSO4 is considered, and the barrier is 44.6 kcal mol−1 (TS1C). This seems difficult to be overcome at room temperature, and thus we omit the discussion on this pathway. 3B undergoes hydrogen atom transfer dehydration, converting POV-based V5 clusters into peroxovanadate intermediates 4viaTS2B. The Oh2–H1 distance decreases from 3.60 Å in 3B to 1.14 Å in TS2B, and Oh1–H1 is elongated from 1.00 Å in 3B to 1.23 Å in TS2B. The ΔG°‡ is 10.6 kcal mol−1 and the ΔG° value is −4.6 kcal mol−1. The rebound of the two V–Ob easily occurs from 4 to peroxovanadate intermediate 5 with a ΔG° value of −5.4 kcal mol−1. Interestingly, this peroxovanadate intermediate is highly active because the barrier of oxygen atom transfer to thioanisole viaTS3 is only 15.6 kcal mol−1 with a high exothermic reaction energy (ΔG° = −41.6 kcal mol−1). The Oh1–S distance decreases from 3.92 Å in 6 to 2.03 Å in TS3 and that of Oh1–V is elongated from 1.79 Å to 1.91 Å in TS3. The results confirm that the generation of peroxovanadate intermediate 5 is the key for oxidation of thioanisole. However, the rate determining step for generation of a peroxovanadate intermediate is the cleavage of the O–O bond with a relatively high activation energy (35.1 kcal mol−1), which is even higher than that (29.1 kcal mol−1) of the direct activation by hydrogen peroxide. Therefore, this pathway is unlikely to proceed, as the activation energy is too high to be overcome at room temperature (experimental conditions). In the next section, another pathway (H–OOH bond cleavage) for generation of the peroxovanadate intermediate needs to be discussed.
Proton transfer can occur at both Ot and Ob sites during H–OOH bond cleavage, as shown in Fig. 4. Starting from intermediate 2A, the H–OOH bond cleavage proceeds viaTS1F, and proton transfer from H2O2 to the Ob atom forms a peroxy-containing intermediate 3F. The Ob–H1 distance decreases from 1.90 Å in 2 to 1.26 Å in TS1F, while the distance of Oh1–V is shortened from 3.66 Å in 2 to 2.12 Å in TS1F, and Oh1–H1 is elongated from 0.98 Å to 1.14 Å in TS1F. The distance of V–Oh1 further decreases to 1.94 Å in 3F accompanied by elongation of the Oh1–H1 bond (2.20 Å). The activation energy of this elementary step is 26.4 kcal mol−1 with an endothermic reaction energy (ΔG° = 23.8 kcal mol−1). The peroxovanadate intermediate is then generated from 3FviaTS2F. The Oh2–H1 distance decreases from 1.66 Å in 3F to 1.06 Å in TS2F, and the Ob–H1 distance is elongated to 1.43 Å in TS2F. The value of ΔG°‡ is 13.3 kcal mol−1, and the ΔG° value is −22.2 kcal mol−1. TS2F is the transition state with the highest energy, so the ΔG°‡ to complete the catalytic cycle is the energy difference between TS2F and 2, which is 37.1 kcal mol−1. On the other hand, the pathway of proton transfer from H2O2 to Ot atoms is also considered, and the barrier is 41.9 kcal mol−1 (TS2D). It is worth noting that rather large ΔG°‡ values for both pathways are also inconsistent with the reaction temperature (298.15 K). The results of these studies indicate that it is not sufficient to consider only the interaction between the {V5O9Cl} cluster and H2O2. However, a comparison of the energy barriers shows that the cleavage of the H–OOH bond is more likely to occur. In this case, the proton is transferred from Oh of H2O2 to the Ob site, not to the Ot site. Therefore, proton transfer plays a crucial role, and solvents and counterions (H2Me2N+) may influence this process, which will be further discussed in the following section.
We further considered the potential influence of electrostatic interactions in the reaction process and introduced two counterions (H2Me2N+) for the simulations. However, as shown in Fig. S12–S14,† for the above reaction pathways, the rate determining step and the important geometric changes of the overall reaction process with the addition of (NH2Me2)+ are almost the same as those without (NH2Me2)+, and there is little difference in activation energy. Consequently, it can be concluded that a negligible influence might be induced by the counterions in the reaction.
The Gibbs activation energy of the catalytic cycle with methanol is shown in Fig. 5. In this case, the methanol molecule primarily influences two stages of proton transfer. Upon the approach of H2O2 and CH3OH to 2-1, the reaction is initiated, resulting in the formation of the adduct 2-2. Intermediate 2-2 is stabilized by –7.0 kcal mol−1 though a triple hydrogen bond: the Ob⋯H1 bond between 2-1 and H2O2 with a distance of 1.90 Å, Oc⋯H2 bond between CH3OH and H2O2 with a distance of 1.69 Å, and a hydrogen bond between H2O2 and H2Me2N+ with a distance of 1.71 Å. From 2-2, the H–OOH bond cleavage proceeds via2-TS1 to form intermediate 2-3, in which proton transfer from H2O2 to the Ob atom is similar to TS1F. It is noteworthy that the distance between Ot2 and hydrogen on the hydroxyl group of methanol (Hc) in 2-TS1 is 1.83 Å, indicating that the methanol molecule forms a hydrogen bonding interaction with the terminal oxygen atom (Ot) on the cluster surface. Simultaneously, the distance between Oc and H1 in 2-3 is 1.62 Å, and the distance between Ot2 and Hc in 2-3 is 1.79 Å. These distances indicate that the methanol molecule, surface oxygen atoms, and hydrogen of H2O2 form a hydrogen bonding network, stabilizing the proton-transfer transition state. The value of ΔG°‡ in this step is 23.3 kcal mol−1, which is moderately lower than that (26.2 kcal mol−1) in the absence of methanol molecule. From 2-3, the second step of proton transfer occurs via2-TS2, where H1 is transferred from Ob to Oh1, resulting in peroxovanadate species. The Ob–H1 distance increases from 1.01 Å in 2-3 to 1.46 Å in 2-TS2, while Oh2–H1 is shortened from 1.54 Å in 2-3 to 1.04 Å in 2-TS2. Furthermore, in 2-TS2, the Ot2–Hc distance is 1.54 Å, and Oc–H2 is 1.16 Å, suggesting a complicated hydrogen bonding network within the structure. Interestingly, the activation energy of this rate-determining step is 25.0 kcal mol−1, which is significantly lower than the 35.8 kcal mol−1 observed in the absence of methanol molecule. Such an activation energy is consistent with the experimental results that reaction occurs at room temperature. The results clearly show that methanol molecules play a significant role in facilitating the proton transfer elementary step. With the assistance of methanol, oxygen atoms on the cluster surface and H2O2 form a hydrogen bonding network that stabilizes the transition state for proton transfer processes for generating peroxovanadate intermediate. Subsequently, the peroxovanadate intermediate 2-5 is formed, with the dissociation of H2O. As discussed above, a rapid oxygen atom transfer from the peroxovanadate intermediate 2-5 to thioanisole then occurs, completing the oxidation process. In this process, the charge transfer from the lone pair (p-orbital) of sulfur to the antibonding orbital (σ*) of the O–O bond is crucial for the oxidation of sulfide. The details are presented in Fig. S16 of the ESI.†
Notably, POV also may undergo a competing H2O2 decomposition like VOSO4, as illustrated in Fig. S17.† However, computational results show that the energy barrier (37.5 kcal mol−1) for H2O2 decomposition is particularly high to be overcome when POVs act as catalysts. In contrast, VOSO4 mediates the same process with a substantially lower barrier of 11.4 kcal mol−1. This side reaction pathway exhibits a relatively lower activation energy compared to sulfide substrate oxidation (23.0 kcal mol−1). The marked catalytic activity of VOSO4 toward H2O2 decomposition consequently diverts a significant fraction of H2O2 from the intended sulfide oxidation process, thereby depressing the overall yield (Fig. 1(c)). Conversely, POV intermediates show negligible catalytic activity for H2O2 decomposition. Therefore, the conversion rate is not always linearly dependent on the barrier of the rate-determining step.
In this work, the transition state of proton transfer is stabilized by the methanol molecule (about 13.4 kcal mol−1). To clearly illustrate the stabilization of the hydrogen-bonding network, the weak interaction analysis of the proton-transfer transition state 2-TS2 was conducted using an independent gradient model based on Hirshfeld partition (IGMH) method, as shown in Fig. 6(b). As shown in Fig. 6(b), there is a large green isosurface between the methyl group of methanol and the oxygen atom on the surface of POV, while a blue isosurface exists between the hydroxyl group of methanol and the hydrogen atom of H2O2 as well as the terminal oxygen atom of POV. These results indicate that CH3OH simultaneously forms hydrogen bonds with POVs (OcHc–Ot2, 1.54 Å) and H2O2 (Oh1H1–Oc, 1.16 Å), promoting proton transfer from POVs to H2O2. Also, a weak interaction exists between the methyl group of methanol and the surface oxygen atoms of POV. This result is consistent with the IRI isosurface maps shown in Fig. 6(c). Therefore, the transition state of the rate-determining step is stabilized by multiple hydrogen bonds.
This process significantly differs from the activation mechanism of H2O2 in the VOSO4-catalyzed system, which involves the cleavage of the H–O bond and the transfer of protons to the bridging oxygen sites of the clusters. In the case of {V5O9Cl}, the vanadium centers are coordinatively saturated. If the reaction is via the radical pathway, one of the V–O bonds needs to be broken to stabilize the radical intermediate. Such a large distortion of {V5O9Cl} will result in significant energy increase (24.5 kcal mol−1 as shown in Fig. S18†), and the pathway is energetically unfavorable (the activation energy is 35.1 kcal mol−1). Meanwhile, the surface of {V5O9Cl} is rich in oxygen atoms, which can form an extensive hydrogen bonding network with H2O2 and solvent methanol molecules, efficiently facilitating the proton transfer and thus driving the protonation pathway. In contrast, the oxygen atoms in VOSO4 cannot form an extensive hydrogen bonding network to stabilize the transition state of proton transfer. Furthermore, the vanadium center is coordinatively unsaturated in the VOSO4-catalyzed system. The available coordination site in VOSO4 can stabilize hydroxyl radicals generated from O–O bond cleavage of H2O2, thus facilitating the radical pathway. In this context, {V5O9Cl} selectively cleaves H–OOH bonds of H2O2 through a proton transfer pathway assisted by multiple hydrogen bonds, while VOSO4 cleaves the HO–OH bond via a hydroxyl radical pathway caused by its the coordinatively saturated V center.
![]() | ||
Fig. 7 Gibbs energy profile (in kcal mol−1) of (NH2Me2)6[(V5O9Cl)3(TDA)6] catalyzed methyl phenyl sulfoxide oxidation. The red dashed line represents the transition state part. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta02513c |
This journal is © The Royal Society of Chemistry 2025 |