Yongqi
Zhuang†
ab,
Yuguo
Dong†
ab,
Yanmei
Zheng
ab,
Wenjun
Zhang
ab,
Lin
Dong
ab and
Zupeng
Chen
*ab
aState Key Laboratory for Development and Utilization of Forest Food Resources, Nanjing Forestry University, Nanjing 210037, China. E-mail: czp@njfu.edu.cn
bJiangsu Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, International Innovation Center for Forest Chemicals and Materials, Nanjing Forestry University, Longpan Road 159, Nanjing 210037, China
First published on 27th November 2025
Lignin, the most abundant natural aromatic polymer on Earth, offers a promising feedstock for the production of value-added aromatic products through selective C–C bond cleavage. However, achieving both high product yield and selectivity under mild conditions, such as under visible-light irradiation, remains a significant challenge. In this study, we address this issue by employing ceria (CeO2) as a support to precisely engineer the dispersion of vanadium oxide through two distinct catalyst pretreatment methods: impregnation–calcination (Im-V/CeO2) and physical mixing–calcination (Pm-V/CeO2). The results demonstrate that while CeO2 alone yields 52% and Pm-V/CeO2 yields 45%, the Im-V/CeO2 catalyst exhibits superior performance, converting 1,2-diphenylethanol to benzaldehyde with an impressive 89% yield and 98% selectivity. This enhanced performance of Im-V/CeO2 is attributed to the introduction of low-polymerized vanadium oxide (VOx) species. Unlike highly polymeric V2O5 crystals, these species facilitate the separation of photogenerated electron–hole pairs and reduce charge transfer resistance, thereby improving photocatalytic efficiency. Additionally, the polar nature of VOx species enhances the activation of lattice oxygen, creating more sites for oxygen adsorption and activation. In-depth mechanistic studies reveal that superoxide radicals play a dominant role in the C–C bond cleavage process. This innovative strategy for achieving high-efficiency photocatalytic C–C bond cleavage through the controlled dispersion of vanadium oxide on a cerium oxide support offers valuable insights for the sustainable valorization of lignin.
Green foundation1. Our work develops cerium oxide nanorod photocatalysts modified with low-polymerized VOx species via a simple impregnation–calcination method, enabling visible-light-driven valorization of lignin β-1 model compounds and analogues under mild conditions, which proceeds without the need for hazardous solvents, external oxidants, chemical additives, or UV-light irradiation.2. The anchored low-polymerized VOx not only enhances the intrinsic oxidative activity of CeO2 but also extends vanadium-catalyzed reactions from homogeneous to heterogeneous systems, eliminating the need for ligands and reducing cost. 3. The catalyst operates efficiently with a vanadium loading of only 2.6 wt%, achieving 90% conversion of the β-1 model substrate with 98% selectivity toward benzaldehyde. Moreover, it demonstrates broad applicability to vicinal diols, affording high yields of aromatic monomers. These advantages establish the promise of such cerium–vanadium catalysts for biomass valorization and the selective synthesis of aromatic fine chemicals. |
Lignin, which accounts for approximately 30% of the weight of lignocellulosic biomass, serves as the largest renewable source of aromatic compounds.6 As a highly irregular polymer, lignin mainly consists of three phenylpropanoid subunits—p hydroxyphenyl (H), guaiacyl (G), and syringyl (S)—interconnected by C–O and C–C linkages.7–9 The efficient valorization of lignin into valuable aromatic chemicals necessitates the cleavage of these bonds while preserving the aromatic rings.10,11 Notably, over 60% of these interunit linkages are C–O bonds, which are more readily cleaved due to their lower bond dissociation energies (217.6–334.7 kJ mol−1) compared to C–C bonds(288.7–514.6 kJ mol−1).12 While this reactivity disparity has historically focused research on C–O cleavage, conventional delignification processes often promote condensation reactions that form additional recalcitrant C–C bonds.13–16 Consequently, the cleavage of these stable C–C linkages becomes an indispensable, yet challenging, prerequisite for efficient lignin utilization. Catalytic lignin oxidation has thus emerged as a promising strategy to address this challenge by targeting the selective cleavage of interunit linkages, particularly the thermodynamically demanding C–C bonds.17 In current research, the β-O-4 linkages have attracted widespread attention due to their high abundance (50–60%) in lignin and the coexistence of both C–C and C–O bonds within the structure. In addition to β-O-4 linkages, lignin also contains other C–C bond types, such as β-1, β-β, and 5-5 linkages.18–21 However, owing to their relatively low natural abundance and high bond dissociation energies, strategies for cleaving these C–C bonds remain considerably less explored.
The β-1 linkage, a typical C–C bond connecting phenylpropane units in lignin, has garnered interest despite its low natural abundance (1–15%) due to its potential as a cleavage target for valorization.22 Current thermal catalytic strategies primarily employ homogeneous vanadium (V) or copper (Cu) complexes, which facilitate oxidative cleavage at ca. 100 °C via substrate coordination.23–25 However, these systems necessitate expensive oxidants, harsh additives, and exhibit high sensitivity to ligand design and reaction conditions, resulting in high operational costs and poor industrial viability.11,22 Consequently, photocatalysis has emerged as a promising platform for C–C bond cleavage under mild conditions. The mechanism typically involves sequential hydroxyl oxidation, dehydration, and C–C bond cleavage. A critical challenge herein is the overoxidation of the hydroxyl group to a ketone, which significantly increases the Cα–Cβ bond strength and severely impedes selective cleavage.26,27 Although homogeneous vanadium photocatalysts, as reported by Liu et al., achieve a high benzaldehyde yield of 86% by utilizing the strong oxidizing capability of high-valent vanadium and coordination with the hydroxyl group of the substrate to form a photoactive complex, they still suffer from metal leaching and low quantum efficiency.8,11 Despite the recognized efficacy of homogeneous vanadium complexes in C–C bond cleavage, this knowledge has not been translated into heterogeneous systems; the development of efficient photocatalytic protocols using common vanadium oxide species (e.g., VOx or V2O5) remains largely unexplored.28,29
In contrast to homogeneous systems, heterogeneous catalysts such as titanium oxide (TiO2) and graphitic carbon nitride (g-C3N4) offer superior structural stability.30–34 However, they often lack the requisite selectivity to suppress undesired overoxidation. For instance, Hou et al. reported that engineering TiO2 nanotubes with CuOx and CeO2 clusters could achieve near-complete conversion with 98% selectivity toward benzaldehyde through precise control over the distribution of CuOx species.6 Nevertheless, this approach requires precise control over support pretreatment and metal loading ratios, which imparts substantial process complexity and cost, thereby hindering scalability. Therefore, a compelling need persists for the development of simple, low-cost, and scalable synthetic strategies to fabricate heterogeneous photocatalysts that simultaneously deliver high selectivity and efficiency for β-1 C–C bond cleavage.
Herein, we report a CeO2-based photocatalyst (Im-V/CeO2) decorated with low-polymerized vanadium oxide (VOx) species and enriched with oxygen vacancies (Ov), synthesized via a simple impregnation–calcination method. This catalyst enables highly selective photocatalytic oxidative cleavage of the C–C bond in 1,2-diphenylethanol to benzaldehyde at room temperature. To elucidate the structure-performance relationship, systematic characterization reveals that the pretreatment method engineers the dispersion state of vanadium oxide species from low-polymerized to crystalline forms, and enables detailed structural analysis. The significantly enhanced catalytic activity is primarily due to an increased oxygen vacancy concentration, resulting from the stronger interaction between low-polymerized VOx (with highly polar V
O bonds) and the CeO2 support compared to crystalline V2O5. This electronic modulation further improves charge separation and reduces charge transfer resistance, as supported by photoelectrochemical measurements. Moreover, mechanistic studies using electron paramagnetic resonance (EPR), radical trapping, and substrate expansion have successfully identified the superoxide radical (˙O2−) as the key active species responsible for Cα–Cβ bond cleavage. Based on the reaction product distribution and previous mechanistic reports, a coherent reaction mechanism is proposed. We believe that the present work can provide insights for the design of simpler and more rational photocatalytic systems, particularly through surface modification engineering, since current strategies in this field predominantly rely on the construction of heterojunctions.
To prepare the octahedral (CeO2-O) CeO2 nanostructure, 1.716 g of Ce(NO3)3·6H2O and 0.015 g of Na3PO4 were dissolved in 20 mL and 140 mL of deionized water, respectively, under vigorous stirring. The Ce(NO3)3·6H2O solution was added dropwise to the Na3PO4 solution and stirred at room temperature for 30 min. The resulting suspension was then transferred to a Teflon lined autoclave and heated at 170 °C for 12 h. The product was recovered by centrifugation, washed with deionized water, dried at 80 °C, and subsequently calcined in air at 600 °C for 2 h, following the same protocol used for CeO2-R and CeO2-C.
The conversion (X) of substrate 1 and the yields (Y) of products were calculated based on the following equations:
| X1 (%) = (n1 reacted × 100%)/n1 added |
| Y1a(1c) (%) = (n1a(1c) produced × 100%)/(n1 added × 2) |
| Y1b(1d) (%) = (n1b(1d) produced × 100%)/n1 added |
To visually characterize the dispersion state of vanadium on the CeO2 surface, high-resolution transmission electron microscopy (HRTEM) was conducted. As shown in Fig. 1b, c and Fig. S1, the representative HRTEM images of Im-V/CeO2 confirmed that the rod-like morphology was retained after impregnation and calcination. Meanwhile, the observed lattice spacing of 3.12 Å corresponds to the CeO2 (111) plane. Notably, the absence of V2O5 lattice fringes—which aligns with the XRD evidence indicating the absence of crystalline vanadia phases—indicates that vanadium species exist in a highly dispersed or amorphous state.38 Furthermore, the high-angle annular dark-field scanning TEM (HAADF-STEM) coupled with the corresponding energy dispersive X-ray (EDX) element mappings (Fig. 1d) further verified the uniform distribution of V, Ce, and O throughout Im-V/CeO2. For Pm-V/CeO2, HRTEM images (Fig. 1e and f) also showed a preserved rod-like structure. Lattice spacings of 3.12 Å and 1.56 Å correspond to the CeO2 (111) and (222) planes, respectively, while a 2.76 Å lattice fringe—matching the V2O5 (011) plane—confirms the presence of crystalline V2O5.37 Meanwhile, HAADF-STEM and EDX mappings (Fig. 1g) demonstrate that V, Ce, and O were homogeneously distributed in Pm-V/CeO2 as well.
X-ray diffraction (XRD) analysis was conducted to confirm the findings from TEM observations. As shown in Fig. 2a, all CeO2 based catalysts exhibited characteristic fluorite CeO2 diffraction peaks at 28.6°, 33.1°, 47.5°, 56.3°, 59.1°, 69.4°, 76.7°, and 79.1° (PDF #34-0394).6 Notably, the Im-V/CeO2 catalyst does not show any diffraction peaks corresponding to crystalline V2O5. This absence, supported by TEM and EDX mappings (Fig. S2a), demonstrates that vanadium species are highly dispersed on the CeO2 surface without forming crystalline V2O5. In contrast, the Pm-V/CeO2 sample displayed additional peaks at 15.3°, 20.3°, 26.1°, 31.0°, 32.4°, and 34.3° (Fig. S3a), corresponding to orthorhombic V2O5 (PDF #41-1426).39 The peak at 32.4° matches the (011) plane of V2O5 observed in TEM, confirming the formation of crystalline V2O5. These results indicate that the nature of vanadium species varies significantly depending on the precursor treatment: physical mixing and calcination yield detectable crystalline V2O5, whereas impregnation-calcination results in vanadium being in an amorphous or super dispersed form.
![]() | ||
| Fig. 2 (a) XRD patterns, (b) Raman spectra, (c) FT-IR spectra, (d) Ce 3d XPS spectra, (e) V 2p XPS spectra, (f) O 1s XPS spectra of CeO2, Im-V/CeO2, and Pm-V/CeO2. | ||
To further elucidate the nature of vanadium species introduced by impregnation, Raman spectroscopy was performed (Fig. 2b). All CeO2-based catalysts exhibit the characteristic F 2g vibration mode of fluorite CeO2 at 460 cm−1. In the case of Im-V/CeO2, this band remains at 460 cm−1, but its intensity is significantly reduced compared to the pristine CeO2, indicating lattice distortion or defect formation upon vanadium incorporation (Fig. S2b).40,41 For Pm-V/CeO2, the Raman spectrum shows additional modes corresponding to crystalline V2O5, in addition to the CeO2 F2 g vibration mode, confirming the formation of V2O5via physical mixing and calcination.42,43 The distinct Raman signatures between Im-V/CeO2 and Pm-V/CeO2 further confirm the non-crystalline nature of vanadium species in Im-V/CeO2. These findings align with the XRD results, ruling out the presence of crystalline V2O5 in Im-V/CeO2.
To further clarify the structural origin of the amorphous vanadium species in Im-V/CeO2, as summarized in Table S3 and illustrated in Fig. S4, N2 adsorption–desorption isotherms, along with corresponding Brunner–Emmett–Teller (BET) surface areas and pore size distributions, were evaluated for all catalysts. All samples exhibit typical type IV isotherms with an H3 hysteresis loop, indicative of mesoporous structure.44 Compared to the pristine CeO2 support, Im-V/CeO2 shows a 30% decrease in BET surface area, suggesting that the impregnation process facilitates more effective ionic deposition of vanadium species within the CeO2 mesopores, partially blocking pore channels and reducing accessible surface area. In contrast, Pm-V/CeO2 demonstrates negligible changes in surface area and pore size, indicating minimal structural alteration by physical mixing. Therefore, the impregnation process, unlike physical mixing, causes a significant reconstruction of the CeO2 structure, which directly dictates the resulting dispersion morphology of the vanadium species. The surface vanadium atomic density was calculated to be 5.5 V nm−2 (surface density = number of surface V atoms/BET surface area), which aligns with previously reported vanadium surface densities (4.2–7.7 V nm−2) for VOx species loaded on the surface of TiO2 and Al2O3.44–46
To directly distinguish between the VOx and V2O5 crystalline phases, Fourier-transform infrared (FT-IR) spectroscopy was performed (Fig. 2c and S2c). The broad band at 3428 cm−1 corresponds to surface hydroxyl (–OH) vibrations,45 while the peak at 1623 cm−1 corresponds to Ce–O stretching of the fluorite CeO2 lattice.47 According to literature reports, the bands at 1044 cm−1 and 977 cm−1 are assigned to V
O terminal stretching vibrations (characteristic of isolated or low-polymerized VOx) and V–O–V bridging vibrations (typical of crystalline V2O5), respectively.48,49 The Im-V/CeO2 sample exhibits a single, intense V
O band at 1044 cm−1 without observable V–O–V peaks, indicating the predominant formation of highly dispersed VOx species. In contrast, Pm-V/CeO2 displays both the 1044 cm−1 (V
O) and 977 cm−1 (V–O–V) bands, whose intensity ratio and peak positions mirror the spectral features of pristine crystalline V2O5 (Fig. S3b). These observations align with Raman results, confirming the absence of crystalline V2O5 peaks in Im-V/CeO2 and demonstrating that physical mixing leads to crystalline V2O5 aggregates, while impregnation promotes the formation of isolated or low-polymerized VOx species.
To further clarify how the distinct vanadium species (dispersed VOxvs. crystalline V2O5) interact with the CeO2 support and potentially govern their catalytic behavior, X-ray photoelectron spectroscopy (XPS) and oxygen temperature-programmed desorption (O2-TPD), supplemented by in situ XPS, were performed. The Ce 3d spectrum of Im-V/CeO2 (Fig. 2d) reveals ten peaks, which could be assigned to Ce4+ (882.4, 889.0, 898.4, 901.1, 907.7, and 916.7 eV) and Ce3+ (880.5, 885.2, 898.8, and 903.8 eV), respectively.50 Compared to pristine CeO2, the Im-V/CeO2 catalyst exhibits a positive shift in Ce 3d binding energy, indicating electron depletion from the CeO2 lattice. Conversely, its V 2p spectrum (Fig. 2e and S5) shows a negative shift relative to pristine V2O5, suggesting electron transfer from CeO2 to vanadium species.51–53 While a similar trend was observed for Pm-V/CeO2, the interaction strength between the vanadium species and the CeO2 support differs significantly. In the case of Im-V/CeO2, the strong interaction between VOx and CeO2, facilitated by the strongly polarized V
O bond, results in a marked decrease in surface Ce3+ content. This is accompanied by an increased V4+ fraction and the largest positive shift in lattice oxygen binding energy, indicative of promoted lattice oxygen activation and oxygen vacancy generation (Fig. 2d and f).54 In contrast, Pm-V/CeO2 shows negligible changes in Ce3+ content, V4+ fraction, and lattice oxygen binding energy, confirming a weaker interaction between the physically mixed V2O5 and CeO2. These differences in interaction strength directly influence catalytic behavior: the strong VOx–CeO2 coupling enhances lattice oxygen activation and facilitates the generation of more oxygen vacancies, while the weak V2O5–CeO2 interaction leads to inferior catalytic performance. This conclusion is corroborated by our subsequent O2-TPD results.
Building on the O2-TPD results, we perform in situ XPS under an O2 atmosphere to explicitly probe the dynamic evolution involving low-polymerized VOx species, lattice oxygen (OL), and oxygen vacancies (Ov). As shown in the in situ XPS spectra (Fig. 3a and c), the Ce 3d peak shifts downward by 0.6 eV, while the lattice oxygen (OL) peak shifts upward by 0.4 eV compared to normal conditions under an O2 atmosphere in the dark. These opposing shifts indicate the electron redistribution from OL to cerium, which leads to an increase in Ce3+ and Ov concentration by the O 1s spectrum.55 Meanwhile, the V 2p XPS spectrum (Fig. 3b and S7) shows a decrease in V4+ content from 63% to 58%, suggesting an increase in the average vanadium oxidation state of vanadium in the surface VOx species. Compared with the pristine support (Fig. S6 and Table S4), the introduction of VOx species results in a marked weakening of the adsorbed oxygen desorption peak at around 130 °C, with the amount of desorbed O2 drastically reduced from 0.321 mmol g−1 to 0.063 mmol g−1, accompanied by a shift of the peak toward higher temperature. species.54,56 This indicates effective activation and conversion of adsorbed oxygen species, while the VOx sites are oxidized by the adsorbed oxygen, resulting in an overall higher oxidation state. Concurrently, the lattice oxygen desorption peak at high temperatures (ca. 650 °C) intensifies, with the amount of desorbed O2 increasing from 0.009 mmol g−1 for the pristine support to 0.076 mmol g−1, accompanied by a shift to a lower temperature. This confirms the facilitated activation and migration of lattice oxygen, which is consistent with the increased oxygen desorption amount and provides a channel for oxygen vacancy formation.57 Furthermore, light irradiation leads to a further decrease in V4+ content with a concomitant increase in the proportion of Ov.
![]() | ||
| Fig. 3 In situ XPS spectra of (a) Ce 3d, (b) V 2p, and (c) O 1s for the Im-V/CeO2 catalyst under dark and light irradiation conditions in an O2 atmosphere. | ||
In summary, O2-TPD and in situ XPS collectively reveal a dynamic cycle: low-polymerized VOx species adsorb and activate O2 at the interface, leading to their oxidation (decreased V4+), which in turn triggers lattice oxygen activation, electron transfer to Ce (increased Ce3+), and oxygen vacancy formation. Light irradiation markedly enhances this entire process via photogenerated charge carriers, thereby stabilizing additional oxygen vacancies. This mechanism accounts for the superior oxidative activity of the catalyst under simultaneous O2 and light conditions.
![]() | ||
| Fig. 4 (a) UV-Vis DRS, (b) Tauc plots, (c) VB-XPS, (d) bandgap diagrams, (e) transient photocurrent responses, and (f) PL spectra of CeO2, Im-V/CeO2, and Pm-V/CeO2. | ||
The bandgaps (Eg) were determined using Tauc plots (Fig. 4b), revealing values of 2.93 eV for CeO2, 2.47 eV for Im-V/CeO2, and 2.60 eV for Pm-V/CeO2. These findings suggest that Im-V/CeO2 has a narrower band gap, enhancing its potential for visible-light-driven catalytic applications. VB-XPS was employed to determine the valence band positions (EVB-XPS) of the catalysts, which were 1.91, 2.27, and 2.36 eV for CeO2, Im-V/CeO2, and Pm-V/CeO2, respectively (Fig. 4c). These values were converted to the Normal Hydrogen Electrode (NHE) scale using EVB(NHE) = EVB-XPS + Φ − 4.44 (the instrument work function (Φ) is calibrated to 4.2 eV).58 The conduction band potentials (ECB) were further calculated according to EVB = ECB + Eg. Therefore, the energy band structures of the catalysts were shown in Fig. 4d. Both Im-V/CeO2 and Pm-V/CeO2 exhibited positive shifts in their conduction band potentials relative to CeO2, indicating their capability to activate O2 for catalytic reactions. Additionally, the positive shifts in their valence band potentials enhanced their oxidation ability, favoring substrate conversion.6,11
To evaluate the efficiency of the separation of photogenerated electrons and holes, transient photocurrent responses, electrochemical impedance spectroscopy (EIS), steady-state photoluminescence (PL), and time-resolved photoluminescence (TR-PL) were performed. As shown in Fig. 4e, Im-V/CeO2 exhibited the highest photocurrent density under visible light irradiation, demonstrating superior electron–hole separation efficiency. The EIS Nyquist plots revealed the smallest semicircle for Im-V/CeO2, signifying the lowest interfacial charge-transfer resistance, whereas Pm-V/CeO2 presented the largest resistance (Fig. S8a). As shown in the steady-state PL spectra, Im-V/CeO2 displayed the lowest emission intensity compared to CeO2 and Pm-V/CeO2 (Fig. 4f), suggesting markedly enhanced photogenerated carrier transfer efficiency and suppressed electron–hole recombination in Im-V/CeO2 photocatalyst, likely due to the presence of VOx species.59 Further evidenced by TR-PL analysis (Fig. S8b), Im-V/CeO2 (1.80 ns) displayed the shortest average lifetime decay compared to CeO2 (1.98 ns) and Pm-V/CeO2 (3.67 ns), further confirming the highest non-radiative decay transition rate and fastest charge transfer in Im-V/CeO2. Collectively, these results demonstrate that the strongly polarized V
O bonds in VOx species enhance the preferential electron transfer from CeO2 to VOx, thereby improving the separation efficiency of photogenerated electrons and holes. These findings align with the XPS results, which showed more significant shifts in the Ce 3d and V 2p binding energies for Im-V/CeO2, indicating a more robust interaction between VOx and CeO2 compared to the physical mixing approach.
Under identical conditions, Im-V/CeO2 exhibited superior performance, achieving an 89% yield of 1a. In comparison, Pm-V/CeO2 achieved only 45%, and bare CeO2 delivered 52% (Fig. 5a and S11), demonstrating that wet impregnation greatly increases active-site density. Moreover, the performance varied with V loading and catalyst amount. Yield increased up to a V loading of 2.58% but plateaued beyond that (Fig. 5b). The optimal yield was achieved at 5 hours of reaction time, after which slight over-oxidation to benzoic acid occurred (Fig. 5c). Similarly, conversion improved with increasing catalyst loading up to 10 mg (Fig. 5d), though yield plateaued likely due to light attenuation. The reaction atmosphere played a critical role, with O2 providing the highest conversion and yield, followed by air, and N2 showing a marked performance decline (Fig. 5e and S12). This observation underscores the critical role of O2 in facilitating C–C bond cleavage and further suggests highlights the essential role of O2 in facilitating C–C bond cleavage and suggests distinct mechanistic pathways under O2versus N2 atmospheres.8 Further investigations with other metal precursors (e.g., W, Al, Fe, and La) under optimal conditions showed negligible activity (Fig. S10b and c), reinforcing VOx as the key active species driving the catalytic performance. Furthermore, the remarkable enhancement over previously reported catalysts in both benzaldehyde selectivity and formation rate highlighted the efficacy of our catalyst design strategy (Table S5). Beyond activity and selectivity, we further assess the green credentials of the Im-V/CeO2 system by evaluating its E-factor and atom economy.60 Specifically, we further quantify the mass of waste generated during the reaction and determined that the Im-V/CeO2 system exhibits an E-factor of only 0.012 in the oxidative C–C bond cleavage of 1,2-diphenylethanol. As summarized in Fig. S14 and Table S6, this value is significantly lower than those reported for conventional photocatalytic systems, highlighting the superior environmental compatibility of our catalyst. To better assess the green and sustainable nature of this system, we also analyse its atom economy. As illustrated in Fig. S15, suppressing the over-oxidation of benzaldehyde to benzoic acid increases the atom economy from 30% to 47.7%. Furthermore, by inhibiting the initial oxidation step that leads to 1,2-diphenylethanone, the atom economy rises substantially to 92.2%, which aligns well with our initial catalyst optimization strategy. Finally, the stability of Im-V/CeO2 was demonstrated through five consecutive stability tests, during which the catalyst maintained near-constant conversion and yield. Post-reaction XRD and FT-IR characterizations confirmed no structural changes (Fig. 5f and S13), thereby showcasing the excellent stability of the catalyst.
When 1 equivalent of p-benzoquinone (p-BQ), a superoxide radical scavenger, was added, the yield of benzaldehyde (1a) decreased to 7%, while the formation of 1,2-diphenylethanone (1b) increased to 55%. This observation confirms that ˙O2− is the key species responsible for C–C bond cleavage and highlights the competitive interplay between hydroxyl oxidation and C–C bond scission.6,11 Furthermore, when 1 equiv. of ammonium oxalate ((NH4)2C2O4) or potassium persulfate (KPS) was used to trap holes or electrons, respectively, the yield of benzaldehyde decreased to 9% and 26%. This demonstrates that both photogenerated holes and electrons are essential for the reaction, underscoring the importance of charge carriers in the photocatalytic process (Fig. 6a).
To investigate the role of oxygen vacancies in photocatalytic processes, electron paramagnetic resonance (EPR) studies were conducted to quantify oxygen vacancy concentrations.2 As shown in Fig. 5b, Im-V/CeO2 exhibits a significantly stronger signal at g = 2.003 compared to CeO2 and Pm-V/CeO2, which correlates directly with its highest activity. This observation, when compared to V2O5, provides evidence for strong electronic coupling between VOx species and the CeO2 lattice.
The polar V
O bonds interact with lattice oxygen, promoting oxygen vacancy formation and enhancing the catalyst's performance in activating O2. Moreover, in situ EPR in a 5,5-dimethyl-1-pyrroline N-oxide (DMPO)/methanol solution under illumination (Fig. 6c) demonstrated that Im-V/CeO2 generated the most intense ˙O2− signal after 10 min, further confirming the role of oxygen vacancies in enhancing O2 activation. Importantly, this signal intensity was quantitatively correlated with the catalytic activity, showing a clear linear relationship (Fig. S17), which further confirms the critical role of ˙O2− radicals in the reaction. To gain a more comprehensive understanding of radical intermediates, in situ DMPO-EPR experiments were performed (Fig. 6d). These revealed no detectable signals in the dark, but progressively intensified six-line peaks under light irradiation, matching β-carbon-centered radicals. These findings corroborated by isotopic evidence, confirm the involvement of a carbon-centered radical pathway.62–64 Substrate scope studies (Fig. 6e) using analogues of substrate 1 under identical conditions provided further insights into the reaction mechanism. Ketone (1b) and 1,2-diphenylethane (1d) remained completely unreacted, ruling out 1b as an intermediate and reinforcing the essential role of the benzylic hydroxyl group in photocatalytic C–C bond scission. Additionally, substrate 2 exhibited negligible conversion, indicating high selectivity for β-1 models. Substrates 3–5, which bear benzylic hydroxyl groups, were fully converted, with 3 and 4 showing superior performance compared to 5. This suggests that side-chain functionality also plays a role in determining selectivity. Collectively, these findings demonstrate that oxygen vacancies enhance the concentration of reactive sites and improve charge separation on the catalyst surface. This facilitates the synergistic action of ˙O2− radicals, electrons, and holes, enabling efficient C–C bond scission and high yields of benzaldehyde.
Photocatalytic conversion of methanol-extracted poplar and birch lignin catalyzed by Im-V/CeO2 was conducted within 24 h.65 After the reaction, the initially crimson lignin solution in acetonitrile faded markedly (Fig. 7a), suggesting catalyst-induced oxidative cleavage. To verify that this change corresponded to actual depolymerization, we analyzed the liquid products by GC-MS and identified various low-molecular-weight aromatic compounds (Fig. 7b).11 Further evidence was provided by gel permeation chromatography (GPC), which showed a clear decrease in the number-average molecular weight (Mn) of birch lignin from 1832 to 1208 g mol−1 after the reaction, along with peak broadening and a shift to higher elution volumes (Fig. 7c)—collectively indicating polymer chain scission.66 These results demonstrate the efficacy of Im-V/CeO2 in cleaving C–C bonds in native lignin. It should be noted, however, that the theoretical maximum yield of cleavage products remains limited to approximately 1%, due to the low natural abundance of β-1 linkages.67
Based on the experimental results outlined above and previous studies, we propose three possible photocatalytic conversion pathways for the conversion of the β-1 model compound using Im-V/CeO2 as the catalyst (Fig. 8). According to the main product types, the pathways can be categorized into two C–C bond cleavage pathways (paths I and II) and one preliminary oxidation pathway (path III). The low selectivity for ketone byproducts (below 2%) suggests that preliminary oxidation is not the prevailing mechanism. This aligns with existing studies and is further confirmed by the ˙O2− trapping experiments discussed earlier, which indicate that preliminary oxidation and C–C bond cleavage are competitive processes.11 Path I, representing the primary reaction pathway, occurs under light irradiation. Oxygen vacancies on the catalyst surface adsorb and activate O2, which is subsequently reduced to ˙O2− by photogenerated electrons. Simultaneously, photogenerated holes attack the Cβ–H bond to generate a Cβ radical intermediate. These radical captures ˙O2− and protons, forming a six-membered ring transition state. This is followed by intramolecular electron transfer, leading to C–C bond cleavage and the formation of two benzaldehyde molecules. The synergistic participation of photogenerated electrons, holes, oxygen vacancies, and molecular oxygen in this pathway explains the significant decrease in benzaldehyde yield during trapping experiments, thereby confirming path I as the dominant mechanism.
![]() | ||
| Fig. 8 Proposed reaction pathways for the photocatalytic conversion of 1,2-diphenylethanol catalyzed by Im-V/CeO2. | ||
Path II resembles the homogeneous vanadium-catalyzed C–C cleavage mechanism reported by Liu et al., which operates in the absence of oxygen.8,11 In this pathway, the benzylic hydroxyl group of the substrate is activated by holes to form a radical, leading to direct cleavage and the production of benzaldehyde and a benzyl radical. The benzyl radical, unable to undergo further oxidation, couples to form 1,2-diphenylethane. In contrast, under a nitrogen atmosphere, only trace amounts of benzaldehyde and 1,2-diphenylethane are observed, further supporting the conclusion that path II is not a dominant mechanism (Fig. S12). Path III corresponds to a preliminary oxidation process, where the hydroxyl group (–OH) is directly oxidized by holes to yield ketone products. However, the low ketone yield suggests that this pathway is not the primary mode of action. It becomes significantly more active only when ˙O2− is inhibited, indicating that it competes with C–C cleavage rather than dominating it.
Footnote |
| † These authors contributed equally to this work. |
| This journal is © The Royal Society of Chemistry 2026 |