Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Hydroxyl-oxygen vacancy synergy over In2O3–ZrO2 catalysts: mechanistic insights into CO2 hydrogenation to methanol

Xia Li a, Zixia Fengb, Hanjun Lua, Xinlin Hong*a, Guangchao Li *bcde and Shik Chi Edman Tsangb
aCollege of Chemistry and Molecular Sciences, Wuhan University, Wuhan 430072, China. E-mail: hongxl@whu.edu.cn
bDepartment of Applied Biology and Chemical Technology, The Hong Kong Polytechnic University, Hong Kong 999077, China. E-mail: guangchao.li@polyu.edu.hk
cThe Hong Kong Polytechnic University Shenzhen Research Institute, The Hong Kong Polytechnic University, Shenzhen 518057, China
dPolyU-Daya Bay Technology and Innovation Research Institute, The Hong Kong Polytechnic University, Huizhou 516000, China
eResearch Institute of Advanced Manufacturing (RIAM), Research Centre for Resources Engineering towards Carbon Neutrality (RCRE), The Hong Kong Polytechnic University, Hong Kong 999077, China

Received 25th April 2025 , Accepted 3rd June 2025

First published on 3rd June 2025


Abstract

The synergistic interplay between oxygen vacancies (OV) and hydroxyl species in In2O3–ZrO2 catalysts plays a crucial role in steering CO2 hydrogenation pathways, however, the atomic-scale interactions between these features have remained elusive. In this study, we engineered In2O3–ZrO2 solid solutions via ZrO2 aerogel phase modulation and thoroughly elucidated the surface chemistry using advanced experimental techniques, including solid-state NMR, in situ DRIFTS, and adsorption studies. The results demonstrate that three distinct hydroxyl site types on the catalyst's surface (terminal hydroxyls (μ1-OH), bridged hydroxyls (μ2-OH), and triply bridging hydroxyls (μ3-OH)) are in close spatial proximity. Besides, μ2-OH and μ3-OH are particularly susceptible to dihydroxylation, a process that facilitates the generation of OV that serve as anchoring sites for CO2. These hydroxyl-vacancy ensembles effectively promote CO2 activation to carbonate/bicarbonate species, which then undergo selective hydrogenation to methanol via a formate-mediated pathway, thus establishing a self-sustaining catalytic cycle. This work clarifies the cooperative role of vacancy coordination and hydroxyl chemistry in CO2 activation and provides a mechanistic guide for the rational design of bimetallic oxide catalysts for CO2 hydrogenation to methanol.


image file: d5ta03275j-p1.tif

Guangchao Li

Dr Guangchao Li obtained his PhD in physical chemistry in 2021 at the University of Chinese Academy of Sciences (UCAS). He was a visiting PhD student at the University of Oxford (2019–2021). Later, he worked as a postdoctoral researcher at the Hong Kong Polytechnic University (PolyU). Since 2023, he has worked as a research assistant professor in the Department of Applied Biology and Chemical Technology at PolyU. His current research interests focus on the characterization and application of porous materials in energy conversion.

1 Introduction

Since the inception of the “methanol economy” and “liquid sunshine”, CO2 hydrogenation to methanol has emerged as a pivotal technology for fossil fuel substitution strategies.1–4 Bimetallic oxide catalysts, particularly those combining redox-active and structural-stabilizing components, have shown significant advantages in optimizing CO2 activation kinetics and methanol selectivity. These benefits arise from their synergistic interfacial interactions and tailored electronic configurations.5,6 However, the atomistic mechanisms driving these processes, particularly defect-mediated catalysis, remain poorly understood due to the inherent complexity of these systems. This complexity necessitates the use of model catalysts to understand underlying structure–activity relationships. This knowledge deficit motivates focused investigation of the In2O3–ZrO2 system, where interfacial oxygen vacancies emerge as catalytic linchpins linking macroscopic performance to atomic-scale phenomena.

Early studies on pure In2O3 revealed oxygen vacancies as critical for CO2 activation, with the coordination environments governing product selectivity.7 Triple-coordinated vacancies on specific crystal planes (e.g., c-In2O3(111)) promote methanol formation via HCOO* stabilization, whereas double-coordinated vacancies favour CO production through C–O bond cleavage.8 This reactivity highlights that vacancy geometry, not merely concentration, is the primary structural factor determining the selectivity of CO2 hydrogenation pathways.9 Since doping ZrO2 brings additional complexity, there is an ongoing debate about the origin of the reactive oxygen vacancies.10,11 Blum et al. attributed these vacancies to a partial reduction of In2O3 at the In2O3/m-ZrO2 interface.12 While Müller et al. demonstrated that the In-OV-Zr structure, formed by the solvation of In3+ into the ZrO2 lattice, serves as the true reactive centre.13 Furthermore, Gong et al. reported that electron transfer modulates the electronic properties of the active sites, enhancing the selectivity toward methanol in ways that transcend the influence of vacancy concentration alone.14 These findings challenge the simple vacancy-centred model, underscoring the multifaceted nature of catalytic selectivity in these systems.

Further investigations reveal that the surface chemical complexity of In2O3–ZrO2 catalysts arises not only from the dynamic evolution of oxygen vacancies but also from the diversity of surface hydroxyl species.15 Metal oxides inherently host a variety of hydroxyl groups, including terminal hydroxyls (μ1-OH), doubly bridged hydroxyls (μ2-OH), and triply bridged hydroxyls (μ3-OH), whose stability and reactivity are governed by their local coordination environments.16 Under thermal pretreatment and hydrogenation conditions, weakly bound hydroxyl groups are selectively destabilized via dehydration or heterolytic H2 dissociation, thereby revealing distinct OV configurations.17 Moreover, the catalytic behaviour varies significantly among these hydroxyl groups. Previous studies have shown that the In2O3−x(OH)y surface is mainly composed of terminally bonded In–OH⋯In and bridged In–OH→In species, with only the former exhibiting the active frustrated Lewis pair (FLP) feature. This FLP efficiently dissociates H2 and directs the selective conversion of CO2 into CO.18–20 This finding suggests that different hydroxyl-vacancy combinations exhibit conformational specificity in regulating the reaction pathways. Nevertheless, there is still a gap in the understanding of the mechanistic details within the more complex In2O3–ZrO2 solid solution system.

In this work, we synthesized a series of composite supports containing monoclinic zirconia and tetragonal zirconia phases by heat treating ZrO2 aerogel precursor. Subsequently, In2O3 was doped onto these supports to obtain In2O3–ZrO2 catalysts featuring diverse hydroxyl sites. By employing magic-angle spinning solid-state nuclear magnetic resonance (MAS SSNMR), in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS), and chemical adsorption analysis techniques, we thoroughly investigated the surface chemical environment of the In2O3–ZrO2 catalysts and elucidated the synergistic interaction between hydroxyl groups and oxygen vacancies in the activation of CO2 molecules. Our findings revealed that double and triple-coordinated OV, especially those in the In–OV–Zr configuration, serve as critical sites for CO2 adsorption and activation. Moreover, hydroxyl groups adjacent to these asymmetric oxygen vacancies further stabilized CO2 and facilitated subsequent hydrogenation processes. These insights provide an understanding of the reaction mechanism underlying CO2 hydrogenation over bimetallic oxide catalysts.

2 Results and discussion

2.1 Catalytic performance for CO2 hydrogenation

Mixed-crystal supports with varying proportions of m-ZrO2 and t-ZrO2 were prepared by calcining amorphous zirconia at 300, 400, 600, and 800 °C. Afterwards, four catalysts, designed as 10In400Zr300, 10In300Zr400, 10In300Zr600, and 10In300Zr800, were synthesized by wet impregnation with indium contents confirmed by ICP-OES (see full description in the Experimental section and Table S1, ESI). It is noteworthy that the relative content of m-ZrO2 in the catalyst increased with increasing calcination temperature of the supports (Table S1). The CO2 hydrogenation performance for these catalysts was evaluated over a temperature range of 250–310 °C at 5 MPa, enabling direct comparison with previous reports on In2O3–ZrO2 catalysts. Prior to testing, all the catalysts were reduced at 350 °C in 10% H2/Ar for 1 h unless otherwise noted. No methanol is formed from pure ZrO2, and the space-time yield (STYMethanol) of bare In2O3 is only 104.3 mg gcat−1 h−1, which is significantly lower than that of 10In300Zr400 (302 mg gcat−1 h−1) at 300 °C, 5 MPa, 12[thin space (1/6-em)]000 mL gcat−1 h−1 (Fig. S1 and Table S2). Additionally, the In2O3–ZrO2 catalyst synthesized via the one-step hydrothermal method performs worse than 10In300Zr400. This may be mostly In doped in ZrO2, which is detrimental to methanol synthesis (Fig. S2 and S5). Due to the competitive relationship between methanol synthesis and the reverse water-gas shift reaction, CO2 conversion and methanol selectivity show a seesaw effect with increasing temperature (Fig. 1a).21 Whereas, the process of increasing temperature promotes the activation of reactants and intermediates and exhibits higher STYMethanol (Fig. 1d). Moreover, the methanol yield is almost unchanged for up to 106 h, which is indicative of exceptional catalyst stability under prolonged reaction conditions (Fig. 1e).
image file: d5ta03275j-f1.tif
Fig. 1 Catalytic performance of 10In400Zr300, 10In300Zr400, 10In300Zr600, and 10In300Zr800 in the CO2 hydrogenation reaction. (a) The correlation between methanol selectivity and CO2 conversion at different reaction temperatures (for the same catalyst, each data point represents a temperature, 250, 270, 290 and 310 °C in that order), (b) CO2 conversion and product distribution, (c) Arrhenius plots for the methanol synthesis reaction kinetics and the corresponding activation energy (Ea) calculations by using Arrhenius equation. (d) Relationship between the space-time yield (STY) of methanol and the reaction temperature. (e) Long-term TOS testing of 10In300Zr400. Reaction conditions: 0.1 g catalyst, GHSV = 12[thin space (1/6-em)]000 mL gcat−1 h−1, 5 MPa, 310 °C, H2/CO2 = 3.

Fig. 1b illustrates the variability in catalytic activity among the four catalysts examined at 290 °C. The results indicate that the catalyst labelled 10In300Zr800 exhibits almost no activity. In contrast, 10In300Zr400 shows superior catalytic performance at this temperature, achieving 83% methanol selectivity and 8% CO2 conversion. Interestingly, as the m-ZrO2 content in the support increases (Table S1), the selectivity of the In2O3–ZrO2 catalyst correspondingly improves. This enhancement in selectivity may be attributed to increased electron transfer between In2O3 and ZrO2. To further investigate the relationship between methanol selectivity and m-ZrO2 content, an identical mass fraction of indium was doped onto both m-ZrO2 and t-ZrO2 supports. Subsequently, the heterogeneous support prepared via the one-pot method was simulated by physical mixing to facilitate a comparative analysis (Fig. S2 and S3). Testing under identical conditions confirmed that methanol selectivity positively correlates with the m-ZrO2 content. Furthermore, the apparent activation energy (Ea) of the methanol synthesis reaction was calculated using the Arrhenius equation (Fig. 1c). The result indicates that the apparent activation energy of the 10In300Zr400 (38.6 kJ mol−1) is noticeably lower than that of 10In400Zr300 (58.9 kJ mol−1), 10In300Zr600 (46.4 kJ mol−1), and 10In300Zr800 (70.2 kJ mol−1), also showing a similar trend as CO2 conversion. Thus, the above results indicate that 10In300Zr400 is most effective at lowering the energy barrier of the methanol synthesis and driving CO2 into methanol.

2.2 Catalyst structure characterization

It is well known that the crystalline phase of ZrO2 can undergo transformation in response to changes in the external environment, particularly temperature fluctuations, as evidenced by TG-DSC, XRD and Raman profiles (Fig. S4–S8).14,22 Previous studies have demonstrated that surface hydroxyl condensation drives the phase evolution of ZrO2. The strong peak at 355 °C observed in the DSC curves is directly associated with the condensation of surface hydroxyls, which facilitates the transformation of a-ZrO2 to t-ZrO2.23,24 In contrast, the weak peak at 481 °C corresponds to the phase change from t-ZrO2 to m-ZrO2.25 Raman spectroscopy with 532 nm excitation was performed on four In2O3–ZrO2 samples. As shown in Fig. S8, the appearance of Raman peaks at 146 and 269 cm−1 correspond to the bulk of t-ZrO2, whereas the m-ZrO2 exhibits distinct Raman peaks at 178, 192, 333, and 378 cm−1.22 The 10In400Zr300 sample displays a dominant tetragonal phase. Notably, the intensity of the monoclinic phase (178 cm−1) increases with higher calcination temperatures, reflecting a gradual phase transition from t-ZrO2 to m-ZrO2. These spectral trends align with the XRD result, corroborating the structural evolution of the ZrO2 support. Moreover, upon doping with 10 wt% In, the diffraction peaks corresponding to the (101) crystal plane (at ∼30.2°, PDF#88-1007) of both fresh and spent 10In300Zr400 shift towards higher angles relative to those of Zr400. This shift indicates the incorporation of In into the ZrO2 lattice, resulting in the formation of InZrOx solid solution (Fig. 2a).13,26 Based on the Scherrer equation, the calculated particle sizes of 10In300Zr400 and 10In400Zr300 are in the range of 9–10 nm (Table S2). The observed difference in catalytic activity between the two catalysts indicates that the crystal size has a relatively minor effect on the overall performance.
image file: d5ta03275j-f2.tif
Fig. 2 Structural characterization and electronic effects. (a) XRD patterns of Zr400, fresh-10In300Zr400 and spent-10In300Zr400. (b) EDX elemental mapping and (c) HAADF-STEM images of fresh-10In300Zr400. (d) H2-TPR profiles of the In2O3–ZrO2 catalysts. XPS profiles of (e) In3d and (f) Zr3d for reduced 10In300Zr400.

Although In2O3 diffraction signals are obscured due to peak overlap with t-ZrO2, aberration-corrected HAADF-STEM images enable clear visualization of its distribution and structure. On 10In300Zr400 catalyst, In2O3 is well dispersed, likely facilitated by the abundant surface hydroxyl groups (Fig. 2b, c and S9).15,27 Images acquired at higher magnification of 10In300Zr400 reveal that the In atom is positioned at the top of the Zr atomic column in the form of brighter spots (Figure S10). This observation aligns with the XRD results, confirming that some of the indium is embedded in zirconium oxide in the form of single atoms.28,29 Moreover, the existence of indium single atoms is still observed on the spent catalysts, suggesting that this structure was extremely stable under high reaction temperatures.

Additionally, H2-TPR measurements were performed to evaluate the H2 dissociation behaviour, which further elucidated the In2O3 morphology across the four In2O3–ZrO2 catalysts (Fig. 2d and 5d). Compared with pure In2O3, the reduction peaks of the In2O3–ZrO2 catalysts shift to higher temperatures, indicating a strong metal–support interaction (SMSI) between ZrO2 and In2O3. This interaction promotes the partial reduction of In2O3, resulting in a more stable In–O–Zr structure and smaller In2O3 particles.30 Also, the higher the proportion of m-ZrO2 in the carrier, the lower the surface reduction temperature of In2O3. XPS analysis of the In 3d and Zr 3d spectra further supports these findings. Previous studies have shown that t-ZrO2 and m-ZrO2 have different electron transfer ability to In2O3, where the interfacial electron transfer from m-ZrO2 to In–O–In is stronger.14,31 In this study, we also observed that the binding energy of In exhibits a shift toward lower values, whereas the binding energy of Zr demonstrates a shift toward to higher values with increasing m-ZrO2 content (Fig. 2e and f). This change suggests that the high electron density of In2O3 may contribute to the methanol selectivity. Moreover, although the H2 dissociation ability of 10In300Zr400 is not the highest among the samples, its H2 consumption is considerably greater than that of other systems, which promotes the formation of oxygen vacancies (Table S3). These oxygen vacancies are essential for the effective activation and dissociation of H2 and CO2, as will be discussed in detail below.

2.3 Identification of surface hydroxyls and oxygen vacancies

The spatial configuration of the surface hydroxyls, which act as a key bridge between the catalyst crystal and the oxygen vacancies, directly affects the exposure of the active sites. 1H MAS SSNMR experiments were conducted to bring out the detailed structural information on the surface hydroxyls of In2O3, different crystalline ZrO2, and the corresponding In2O3–ZrO2 catalysts. As illustrated in Fig. 3a and S11, In2O3, ZrO2, and In2O3–ZrO2 catalysts possess three types of surface hydroxyls. Within the range of 0–10 ppm, characteristic peaks of the surface hydroxyls emerge, with chemical shifts at 0–1.5 ppm, 1.6–3.8 ppm, and 3.9–5.0 ppm corresponding to the terminal hydroxyls (μ1-OH), doubly bridged hydroxyls (μ2-OH), and triply bridged hydroxyls (μ3-OH), respectively (Fig. 3b).16,32 The signals in the high-frequency regions (5.1–8.5 ppm) are attributed to the overlapping resonance of μ2-OH and μ3-OH, which contribute to the formation of extensive hydrogen-bonding networks. This spatial proximity is corroborated by the 2D 1H–1H homonuclear correlation SSNMR spectra (Fig. 3c, S13–S15), where strong diagonal peaks reveal the mutual interactions among μ1-OH, μ2-OH, and μ3-OH. Additionally, both In2O3 and ZrO2 exhibit a μ2-OH characteristic peak at 3.2 ppm, however, in the In2O3–ZrO2 system, a new peak appears at 2.0 ppm. This observation suggests a significant alteration in the hydroxyl coordination environment, attributable to the formation of an In–O(H)–Zr heterointerface. Combining XRD and aberration-corrected HAADF-STEM results, it is inferred that the dissolution of In species onto the ZrO2 surface leads to a partial reorganization of μ3-OH groups. This reorganization results in the formation of an In–O(H)–Zr structure characterized by bimetallic coordination, directly contributing to a reduction in the μ3-OH content in the composite catalyst and an increase in the ratio of μ2-OH groups. These findings confirm the critical role of surface hydroxyls in stabilizing the active indium species and in determining their spatial distribution.
image file: d5ta03275j-f3.tif
Fig. 3 Characterization of surface hydroxyl configurations and their structure–activity correlations. (a) 1H MAS SSNMR spectra of In2O3–ZrO2 catalysts and ZrO2 supports, (b) schematic representation of the three types of surface hydroxyls, (c) 2D 1H–1H homonuclear correlation SSNMR spectra of 10In300Zr400, (d) NH3-TPD profiles of In2O3–ZrO2 catalysts, (e) the proportion of μ2-OH and μ3-OH total acidity and methanol yield for CO2 hydrogenation over In2O3–ZrO2 catalysts, (f) 1H MAS SSNMR spectra of H2-D2 exchange experiment for 10In300Zr400. Reaction condition: 0.1 g catalyst, GHSV = 12[thin space (1/6-em)]000 mL gcat−1 h−1, 5 MPa, 310 °C, H2/CO2 = 3.

By analysing the ratio of various hydroxyl groups in 1H MAS SSNMR and the total amount of acid sites obtained from NH3-TPD measurements, we accurately calculated the specific concentrations of different hydroxyl species in the catalysts. Upon analysing the NH3-TPD profile through peak fitting (as detailed in Fig. 3d and Table S4), a volcanic-like relationship emerges between the total acidity of the In2O3–ZrO2 catalysts and their monoclinic phase content, with the catalyst labelled 10In300Zr400 possessing the highest concentration of acid sites (182 μmol g−1). When correlating the total acidity of the catalysts with the methanol yield, a positive correlation trend akin to that between oxygen vacancy concentration and methanol yield is observed, which points to a link between oxygen vacancies and surface hydroxyls (Fig. S16). Intriguingly, the total μ2-OH and μ3-OH concentrations of four In2O3–ZrO2 catalysts show regular variations: 10In300Zr400 > 10In300Zr600 > 10In400Zr300 > 10In300Zr800, and this order coincides with a significant linear correlation with their methanol yield (Fig. 3e). Taking this into account, it is reasonable to speculate that μ2-OH and μ3-OH may be the active site for methanol synthesis. To confirm this point, we adsorbed H2 on the catalyst surface and then switched D2 to exchange the surface hydrogen species to obtain the 1H SSNMR spectra (Fig. 3f). Our results show that a significant attenuation of both μ2-OH and μ3-OH signals in the 2.0–8.5 ppm region after D2 treatment suggests that the two types of hydroxyl groups are highly reactive. In corroboration with this, O 1s XPS analysis shows a decrease in the hydroxyl group fraction after the reduction treatment and an increase in the oxygen vacancy fraction. Therefore, we propose that μ2-OH and μ3-OH are structural precursors of the active site, which may be generated by dehydroxylation to oxygen vacancies during the reaction process.

More information on the density of oxygen vacancies is unravelled in the O 1s spectra by XPS. As shown in Fig. 4a, the broad peak of the O 1s spectra can be deconvoluted into three different components. The peak at the lower binding energy (529.8 eV) is associated with the lattice oxygen of ZrO2 (OL). The two peaks centered at ca. 531.2 eV and 532.4 eV originate from O atoms in the proximity of an oxygen vacancy (OV) and the chemisorbed oxygen species related to the surface hydroxyls (*OH). The radio of OV/(OL + OV + OH) and OH/(OL + OV + OH) reflects the density of surface oxygen vacancies and the surface hydroxyls, respectively.30,33 It is worth noting that doping In on Zr400 increases the concentration of OV without changing *OH concentration, regarded as interfacial oxygen vacancies formed via the penetration of In into ZrO2. Also, after reducing 10In300Zr400 by H2, the oxygen vacancy concentration is increased from 21.4% to 26.9%, while the hydroxyl group concentration is decreased from 11.3% to 9.26% (Fig. S17). Thus, some of the oxygen vacancies on the catalyst surface are derived from the oxidation of surface hydroxyls. Surface oxygen vacancies, as electron-rich centres, can effectively activate CO2 molecules, and are usually regarded as important active sites for methanol synthesis by CO2 hydrogenation. Systematic comparison of the composition–performance relationships among different catalyst systems revealed that OV and OH concentrations in reduced catalysts follow the order 10In300Zr400 > 10In300Zr600 > 10In400Zr300 > 10In300Zr800, both exhibiting linear correlations with methanol synthesis activity (Fig. 4b and S18). This result aligns with Liu et al.'s study,15 where air plasma treatment enhanced the hydrophilicity of ZrO2. In the highly hydrophilic 1% Ru/DBD-ZrO2 catalyst, CO2 is captured at oxygen vacancies, and the H atoms from surface OH* combine with CO2 to form HCO3* intermediates, which further convert to HCOO*. Although their work focused on CO2 hydrogenation to CO, we propose that oxygen vacancies and hydroxyl groups serve as critical active sites and descriptors for catalyst structure–activity relationships, given the established role of formate as a key intermediate in methanol synthesis.


image file: d5ta03275j-f4.tif
Fig. 4 Characterization for revealing the relationship between oxygen vacancies and CO2 activation. (a) Deconvoluted O 1s XPS spectra In2O3–ZrO2 catalysts, (b) the proportion of Ov concentration and methanol yield for CO2 hydrogenation over In2O3–ZrO2 catalysts, (c) 31P MAS SSNMR spectra of TMP adsorbed on In2O3–ZrO2 catalysts, (d) CO2-TPD profiles of In2O3–ZrO2 catalysts, (e) 1D 13C MAS NMR spectra of CO2 adsorbed on 10In300Zr400 and 10In300Zr600. (f) 2D 13C–1H heteronuclear correlation MAS SSNMR spectra of 10In300Zr400.

SSNMR is a powerful technique for the characterization of active sites on the solid materials.34–36 As the TMP molecule is highly alkaline, TMP-assisted 31P MAS SSNMR analysis was adopted to distinguish the acid sites over the catalysts.37 Previous studies have shown that TMP is more sensitive to Lewis acidity than Brønsted acidity, and the signal ranging from −30 to −58 ppm is attributed to TMP binding onto the LAS.38 Fig. 4c shows that for all catalysts except 10In300Zr800, two peaks are observed at −50 to −60 ppm and 40 to 60 ppm, respectively. The signals at −60, −57, and −55 ppm are assigned to Lewis acid sites in 10In400Zr300, 10In300Zr600, and 10In300Zr400, respectively. Apparently, 10In300Zr400 exhibits stronger Lewis acidity than both 10In300Zr600 and 10In400Zr300, which aligns with the increased oxygen vacancy concentration observed in the O 1s spectra. The signal in the positive chemical shift range corresponds to TMPO, formed by the reaction between TMP and lattice oxygen, indicating that the In2O3–ZrO2 catalyst readily forms more oxygen vacancies.39

To investigate the effect of Lewis acidity/basicity of In2O3–ZrO2 catalysts on the CO2 adsorption and activation, the CO2-TPD measurement was performed, as CO2 molecules could bind strongly to FLP. The signals of CO2 were collected by the mass spectrometer. As shown in Fig. 4d, CO2 desorption peaks can be divided into three regions, 0–200 °C, 200–400 °C, and 400–900 °C, which are attributed to weak, medium-strong, and strong basic sites, respectively. In the temperature range below 400 °C, the two signals correspond to the physically adsorbed CO2 and CO2 adsorbed on the weakly basic site, like surface hydroxyls.40,41 In the high temperature range of 400–900 °C, three consecutive desorption peaks at around 560, 740, and 900 °C are observed for all catalysts. These peaks are assigned to CO2 adsorbed as bidentate carbonates (b-CO32−) and monodentate carbonates (m-CO32−) on the strongly basic sites (OV).42 Compared with the CO2-TPD profiles of pure ZrO2, we noticed that a larger peak area in the high temperature region for 10In300Zr400, which indicates that the oxygen vacancy concentration is elevated after doping with In, and this result is also evidenced by H2-TPR and O 1s XPS (Fig. S19). Analogously, the existence of more surface oxygen vacancies on 10In300Zr400 is further verified. To explore the effect of H2 pretreatment on the surface properties of the catalysts, the In2O3–ZrO2 catalysts were exposed to H2 and Ar atmospheres at 350 °C for reduction and heat treatment, respectively (Fig. S20). Obviously, after treatment with H2, the signal intensities in the low-temperature region (<200 °C) decrease while the peak in the high-temperature region increases, suggesting that H2 reduction creates more oxygen vacancies by removing the surface hydroxyls.

Meanwhile, high-resolution 1D 13C MAS SSNMR and 2D 13C–1H Heteronuclear Correlation (HETCOR) MAS SSNMR spectra have been utilized to precisely elucidate the adsorption behaviour and structure of CO2 over the catalyst surface. In general, the electron-rich oxygen atoms in the CO2 molecule are anchored by Lewis acids (OV) and are immobilized on the catalyst surface in the form of bidentate carbonates (b-CO32−) and monodentate carbonates (m-CO32−). As depicted in Fig. 4e, three 13C signals are observed in the 158–175 ppm region after introducing 13CO2 to In2O3–ZrO2. The peaks from 158 to 162 ppm, 163 to 167 ppm, and 167 to 171 ppm correspond to bicarbonate, bidentate carbonate, and monodentate carbonate, respectively.43 Furthermore, the 2D 13C–1H HETCOR MAS SSNMR experiments reveal the spatial proximity of the adsorbed 13CO2 and the three surface hydroxyls (Fig. 4f and S21). Apparently, the b-CO32− species in the 163–167 ppm regime exhibits strong interactions with structural hydroxyl groups of μ2-OH and μ3-OH, as well as their associated hydrogen-bonding network. No correlation peak corresponding to an interaction with μ1-OH is observed. Therefore, we conclude that μ2-OH and μ3-OH adjacent to OV act synergistically as proton donors to promote CO2 protonation to HCO3 on the one hand, and by stabilizing b-CO32− through the O–H⋯O bond on the other, thereby providing an optimized microenvironment for CO2 hydrogenation.

2.4 Investigation of the reaction mechanism

In situ DRIFTS measurements were conducted to elucidate the synergistic effect of surface hydroxyls and oxygen vacancies in methanol synthesis. Fig. 5a presents the real-time monitoring of the catalytic process over the 10In300Zr400 catalysts under programmed temperature conditions. When the test temperature is controlled at 100 °C, a broad absorption band in the range of 1660 and 1200 cm−1 is observed, which can be attributed to the formation of carbonate or bicarbonate species. Specifically, the peaks at 1276, 1300, 1662 cm−1 and 1508, 1335 cm−1 are assigned to bidentate carbonate species (b-CO32−) and monodentate carbonate (m-CO32−), respectively. In addition, peaks at 1446 and 1624 cm−1 correspond to bidentate bicarbonate species (b-HCO3).44 These findings suggest that CO2 is predominantly adsorbed on the reduced catalyst as carbonates and bicarbonates, consistent with CO2-TPD and 13C MAS SSNMR results. Moreover, linearly adsorbed CO (2077 cm−1) is also observed at 100 °C, which may originate from the dissociation of carbonates induced by surface oxygen vacancies, as supported by the literature.42 As the temperature increases, the band of carbonate species and bicarbonate species gradually diminish, while new spectral features emerge. These new features correspond to the formation of bidentate formates (b-HCOO*, 1392, 1376, 1592 cm−1) and methoxy (CH3O*, 1046 and 1140 cm−1).45,46 Notably, upon reaching 250 °C, the HCOO* absorption peak intensity markedly increases, while the HCOO*/CH3O* ratio initially rise but subsequently decline, suggesting that HCOO* acts as a key intermediate in the CO2 hydrogenation pathway, with its conversion to CH3O* requiring a high activation energy barrier (Fig. S22). Furthermore, time-resolved DRIFTS analysis at 300 °C captures the dynamic behaviour of reaction intermediates (Fig. 5b). Initially observed carbonate and bicarbonate species, analogous to those at 100 °C, show rapid intensity decay within 10 minutes, while HCOO* and CH3O* species progressively increase, indicating the conversion of carbonate/bicarbonate to formate on the In2O3–ZrO2 catalysts.
image file: d5ta03275j-f5.tif
Fig. 5 In situ DRIFTS spectra of CO2 + H2 reaction in the range of 1000–2200 cm−1 over 10In300Zr400 catalyst. (a) DRIFTS spectra at different temperatures (100–300 °C) after reacting for 1 h, (b) time-resolved DRIFTS spectra at a reaction temperature of 300 °C; (c) DRIFTS spectra of H2 reduction at 350 °C; (d) H2-TPR profiles of ZrO2, In2O3, 10In400Zr300.

To further elucidate the hydrogen dissociation mechanism on the In2O3–ZrO2 catalyst, we also conducted in situ DRIFTS experiments under hydrogen-rich conditions at 350 °C and 1 MPa. As illustrated in Fig. 5c, upon exposure to a hydrogen atmosphere, the characteristic peak corresponding to surface hydroxyls (OH*, 3684 cm−1) on the 10In300Zr400 catalyst demonstrates a progressive intensity reduction. Meanwhile, distinct shifts at 1707, 1584, 1533, and 1390 cm−1 appear, which can be assigned to the intermediate In2H4, Zr–H, InInH2, and InH species.47 Previous studies have demonstrated that molecular hydrogen preferentially adsorbs on the In2O3 surface, where it readily undergoes heterolytic dissociation to form indium hydride species.19 This preferential adsorption behaviour is supported by both theoretical calculations and surface characterization techniques.7,48 In contrast, H2-TPR has revealed that ZrO2 exhibits a strong tendency to interact with carbon-based species rather than hydrogen molecules, indicating its limited capability for hydrogen dissociation under similar conditions (Fig. 5d).49,50 These collectively suggest that while hydrogen molecules remove surface hydroxyls, they simultaneously generate electron-rich InH species on the indium oxide surface. Accordingly, Zr–H species are likely formed through a hydrogen spillover mechanism from the adjacent In2O3 surface, rather than through direct dissociation on ZrO2.11 This hydrogen migration phenomenon is consistent with the well-documented surface mobility of hydrogen atoms in mixed oxide systems.

2.5 Discussion on hydroxyl-oxygen vacancy synergy

Based on the experimental results discussed above, Fig. 6 illustrates the proposed surface chemical environment of the In2O3–ZrO2 catalyst and delineates the synergistic mechanism between hydroxyl groups and oxygen vacancies in the activation of CO2 molecules. The crystalline phase of ZrO2 is usually closely related to the distribution of surface hydroxyls. We found that 400 °C approaches the critical phase transition temperature of ZrO2, and calcination at temperatures above this threshold enables the formation of mixed-crystalline supports. The 10In300Zr400 catalyst constructed on mixed-crystalline ZrO2 by low-temperature calcination is enriched with a higher OV concentration at the interface while simultaneously maintaining a greater density of surface hydroxyls. According to XPS and XRD results, the more severe lattice mismatch between the mixed-crystalline ZrO2 and In2O3, which creates a greater number of interfacial oxygen vacancies than in pure t-ZrO2.28 Compared to the physical mixed system with the same mixed-crystal composition, the catalysts synthesised by the one-pot method exhibit higher CO2 conversion than, confirming that in situ construction of interfacial oxygen vacancies is the key to the formation of active sites (Fig. S2). The atomic-level dispersion of In2O3 is promoted through surface hydroxyls and strong metal–support interactions, thereby increasing the number of active sites. However, further elevation of calcination temperatures leads to particle coarsening and reduced hydroxyl group concentration, resulting in a reduction of hydroxyl-oxygen vacancy active sites and a decline in catalytic performance. During CO2 hydrogenation, H2 preferentially dissociates at In-containing structural sites, such as In nanoclusters and the In–O(H)–Zr structure. This process exposes oxygen vacancies that exist in double- and triple-bridged coordination environments. These oxygen vacancies function as strong Lewis acid sites, anchoring the oxygen atoms of CO2, while the neighbouring μ2-OH/μ3-OH network forms FLP sites through hydrogen bonding. This synergistic interaction promotes the formation of carbonate and bicarbonate in a locally polar environment. The CO2 species stabilized by the hydroxyl network are activated in the presence of the In–H active site, leading to directed hydrogenation to methanol via formate intermediate. Throughout the reaction, the depleted hydroxyl groups and oxygen vacancies are dynamically regenerated by surface remodelling under reaction conditions, thereby creating a hydroxyl-oxygen vacancy synergistic catalytic cycle.
image file: d5ta03275j-f6.tif
Fig. 6 Schematic representation of CO2 hydrogenation to methanol over In2O3–ZrO2 catalysts.

3 Conclusions

In summary, a straightforward In2O3–ZrO2 model catalyst was utilized to investigate the distribution of surface hydroxyls and to elucidate the synergistic interactions and transformations between oxygen vacancies and surface hydroxyls in CO2 hydrogenation to methanol. Among the catalysts tested- 10In400Zr300, 10In300Zr400, 10In300Zr600, and 10In300Zr800 -the 10In300Zr400 catalyst exhibits a higher concentration of surface hydroxyls and oxygen vacancies. Under the operating conditions of 300 °C, 5 MPa, and 12[thin space (1/6-em)]000 mL gcat−1 h−1, it achieved superior methanol synthesis activity (302 mg gcat−1 h−1). Structural characterization and mechanistic studies revealed that the FLP sites formed by surface hydroxyls and oxygen vacancies, act as centres for stabilizing CO2, providing an abundant carbon source for the formation of methanol synthesis intermediates, HCOO* and CH3O*. Overall, this work reinterprets the physicochemical behaviour of small molecule reactants on the catalyst surface from an acid–base perspective, offering novel insights for designing catalysts with high-density active interfacial sites for enhanced performance.

Data availability

The data supporting this article have been included as part of the ESI.

Author contributions

Xia Li: data curation, software, formal analysis, writing–original draft, writing–review & editing. Zixia Feng: investigation, data curation, formal analysis, resources, methodology. Hanjun Lua: investigation, formal analysis. Xinlin Hong: writing– review & editing, guiding the project. Guangchao Li: writing– review & editing, guiding the project. Shik Chi Edman: discussion.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The support from the Hong Kong Polytechnic University (PolyU P0049034, P0055259) and the Department of Science and Technology of Guangdong Province (GDSTC 2025A1515011688) is gratefully acknowledged. G. Li gratefully acknowledges support from the University Research Facility in Chemical and Environmental Analysis (UCEA) at PolyU. We thank Dr Xu Zhang from City University of Hong Kong for providing the images of the samples.

Notes and references

  1. G. A. Olah, Angew. Chem., Int. Ed., 2005, 44, 2636–2639 CrossRef PubMed.
  2. C. F. Shih, T. Zhang, J. Li and C. Bai, Joule, 2018, 2, 1925–1949 CrossRef.
  3. P. Gao, L. Zhang, S. Li, Z. Zhou and Y. Sun, ACS Cent. Sci., 2020, 6, 1657–1670 CrossRef PubMed.
  4. H. Wang, X. Liu, P. Niu, S. Wang, J. Shi and L. Li, Matter, 2020, 2, 1377–1413 CrossRef.
  5. F. Tao, Chem. Soc. Rev., 2012, 41, 7977–7979 RSC.
  6. X. Jiang, X. Nie, X. Guo, C. Song and J. G. Chen, Chem. Rev., 2020, 120, 7984–8034 CrossRef PubMed.
  7. J. Ye, C. Liu, D. Mei and Q. Ge, ACS Catal., 2013, 3, 1296–1306 CrossRef.
  8. S. Dang, B. Qin, Y. Yang, H. Wang, J. Cai, Y. Han, S. Li, P. Gao and Y. Sun, Sci. Adv., 2020, 6, eaaz2060 CrossRef PubMed.
  9. W. Wang, K. Huo, Y. Wang, J. Xie, X. Sun, Y. He, M. Li, J. Liang, X. Gao, G. Yang, S. Lin, F. Cao, H. Jiang, M. Wu and N. Tsubaki, ACS Catal., 2024, 14, 9887–9900 CrossRef.
  10. N. H. M. D. Dostagir, C. R. Tomuschat, K. Oshiro, M. Gao, J.-y. Hasegawa, A. Fukuoka and A. Shrotri, JACS Au, 2024, 4, 1048–1058 CrossRef PubMed.
  11. F. Gao, Y. Wang, Y. Zhao, K. Wang, W. Guo, Z. Sun, Y. Zhu, H. He, Y. Liu and Y. Cao, ACS Catal., 2025, 15, 2785–2795 CrossRef.
  12. X. Zhang, A. V. Kirilin, S. Rozeveld, J. H. Kang, G. Pollefeyt, D. F. Yancey, A. Chojecki, B. Vanchura and M. Blum, ACS Catal., 2022, 12, 3868–3880 CrossRef.
  13. A. Tsoukalou, P. M. Abdala, A. Armutlulu, E. Willinger, A. Fedorov and C. R. Müller, ACS Catal., 2020, 10, 10060–10067 CrossRef.
  14. C. Yang, C. Pei, R. Luo, S. Liu, Y. Wang, Z. Wang, Z.-J. Zhao and J. Gong, J. Am. Chem. Soc., 2020, 142, 19523–19531 CrossRef PubMed.
  15. M. Liu, R. Zou and C.-J. Liu, Appl. Catal., B, 2025, 360, 124549 CrossRef.
  16. Q. Han, P. Gao, L. Liang, K. Chen, A. Dong, Z. Liu, X. Han, Q. Fu and G. Hou, Anal. Chem., 2021, 93, 16769–16778 CrossRef PubMed.
  17. L. Fu, Z. Qu, L. Zhou and Y. Ding, ACS Sustain. Chem. Eng., 2024, 12, 18412–18421 CrossRef.
  18. G. C. Welch, R. R. S. Juan, J. D. Masuda and D. W. Stephan, Science, 2006, 314, 1124–1126 CrossRef PubMed.
  19. K. K. Ghuman, L. B. Hoch, T. E. Wood, C. Mims, C. V. Singh and G. A. Ozin, ACS Catal., 2016, 6, 5764–5770 CrossRef.
  20. R. Jing, X. Lu, J. Wang, J. Xiong, Y. Qiao, R. Zhang and Z. Yu, Small, 2024, 20, 2310926 CrossRef PubMed.
  21. S. Tada, S. Kayamori, T. Honma, H. Kamei, A. Nariyuki, K. Kon, T. Toyao, K.-i. Shimizu and S. Satokawa, ACS Catal., 2018, 8, 7809–7819 CrossRef.
  22. M. Li, Z. Feng, P. Ying, Q. Xin and C. Li, Phys. Chem. Chem. Phys., 2003, 5, 5326–5332 RSC.
  23. X. Guo, D. Mao, G. Lu, S. Wang and G. Wu, Catal. Commun., 2011, 12, 1095–1098 CrossRef.
  24. Y. Wu, J. Chen, W. Hu, K. Zhao, P. Qu, P. Shen, M. Zhao, L. Zhong and Y. Chen, J. Catal., 2019, 377, 565–576 CrossRef.
  25. Z. Yanwei, G. Fagherazzi and S. Polizzi, J. Mater. Sci., 1995, 30, 2153–2158 CrossRef.
  26. T. Yan, N. Li, L. Wang, W. Ran, P. N. Duchesne, L. Wan, N. T. Nguyen, L. Wang, M. Xia and G. A. Ozin, Nat. Commun., 2020, 11, 6095 CrossRef.
  27. H. Li, W. Wang, S. Xue, J. He, C. Liu, G. Gao, S. Di, S. Wang, J. Wang, Z. Yu and L. Li, J. Am. Chem. Soc., 2024, 146, 9124–9133 CrossRef PubMed.
  28. M. S. Frei, C. Mondelli, A. Cesarini, F. Krumeich, R. Hauert, J. A. Stewart, D. Curulla Ferré and J. Pérez-Ramírez, ACS Catal., 2019, 10, 1133–1145 CrossRef.
  29. H. Li, S. Di, P. Niu, S. Wang, J. Wang and L. Li, Energy Environ. Sci., 2022, 15, 1601–1610 RSC.
  30. Y. Zhou, L. Liu, G. Li and C. Hu, ACS Catal., 2021, 11, 7099–7113 CrossRef.
  31. C. Shen, K. Sun, R. Zou, Q. Wu, D. Mei and C. J. Liu, Appl. Catal., B, 2025, 361, 124683 CrossRef.
  32. J. Qin, Y. Long, F. Sun, P. P. Zhou, W. D. Wang, N. Luo and J. Ma, Angew. Chem., Int. Ed., 2021, 61, e202112907 CrossRef PubMed.
  33. J. An, Y. Wang, J. Lu, J. Zhang, Z. Zhang, S. Xu, X. Liu, T. Zhang, M. Gocyla, M. Heggen, R. E. Dunin-Borkowski, P. Fornasiero and F. Wang, J. Am. Chem. Soc., 2018, 140, 4172–4181 CrossRef PubMed.
  34. G. Li, C. Foo, X. Yi, W. Chen, P. Zhao, P. Gao, T. Yoskamtorn, Y. Xiao, S. Day, C. C. Tang, G. Hou, A. Zheng and S. C. E. Tsang, J. Am. Chem. Soc., 2021, 143, 8761–8771 CrossRef PubMed.
  35. G. Li, T. Yoskamtorn, W. Chen, C. Foo, J. Zheng, C. Tang, S. Day, A. Zheng, M. M. J. Li and S. C. E. Tsang, Angew. Chem., Int. Ed., 2022, 61, e202204500 CrossRef PubMed.
  36. G. Li, C. Foo, R. Fan, M. Zheng, Q. Wang, Y. Chu, J. Li, S. Day, P. Steadman, C. Tang, T. W. B. Lo, F. Deng and S. C. E. Tsang, Science, 2025, 387, 388–393 CrossRef CAS PubMed.
  37. M. Wang, L. Zheng, G. Wang, J. Cui, G. L. Guan, Y.-T. Miao, J.-F. Wu, P. Gao, F. Yang, Y. Ling, X. Luo, Q. Zhang, G. Fu, K. Cheng and Y. Wang, J. Am. Chem. Soc., 2024, 146, 14528–14538 CrossRef CAS PubMed.
  38. D. Rivera-Barrera and J. C. Poveda-Jaramillo, J. Solid State Chem., 2021, 294, 121862 CrossRef CAS.
  39. Z. Zhang, J. Wu, Z. Feng, X. Xia, J. Wen, F. Wang, X. Ning, Y. Wen, X. Ke, H. Guan and L. Peng, J. Phys. Chem. C, 2024, 129, 508–517 CrossRef.
  40. S. Lu, H. Yang, Z. Zhou, L. Zhong, S. Li, P. Gao and Y. Sun, Chin. J. Catal., 2021, 42, 2038–2048 CrossRef CAS.
  41. R. Zou, C. Shen, K. Sun, X. Ma, Z. Li, M. Li and C.-J. Liu, J. Energy Chem., 2024, 93, 135–145 CrossRef CAS.
  42. C. Zhu, X. Wei, W. Li, Y. Pu, J. Sun, K. Tang, H. Wan, C. Ge, W. Zou and L. Dong, ACS Sustain. Chem. Eng., 2020, 8, 14397–14406 CrossRef CAS.
  43. Y. Fu, L. Zhang, B. Yue, X. Chen and H. He, J. Phys. Chem. C, 2018, 122, 24094–24102 CrossRef.
  44. C. Y. Regalado Vera, N. Manavi, Z. Zhou, L.-C. Wang, W. Diao, S. Karakalos, B. Liu, K. J. Stowers, M. Zhou, H. Luo and D. Ding, Chem. Eng. J., 2021, 426, 131767 CrossRef.
  45. L. Liu, F. Cannizzaro, A. Kaychouhi, N. Kosinov and E. J. M. Hensen, Chem. Eng. J., 2024, 494, 153204 CrossRef.
  46. A. R. Richard and M. Fan, ACS Catal., 2017, 7, 5679–5692 CrossRef CAS.
  47. L. Andrews and X. Wang, Angew. Chem., 2004, 116, 1738–1741 CrossRef.
  48. Y. Dong, K. K. Ghuman, R. Popescu, P. N. Duchesne, W. Zhou, J. Y. Y. Loh, F. M. Ali, J. Jia, D. Wang, X. Mu, C. Kübel, L. Wang, L. He, M. Ghoussoub, Q. Wang, T. E. Wood, L. M. Reyes, P. Zhang, N. P. Kherani, C. V. Singh and G. A. Ozin, Advanced Science, 2018, 5, 1700732 CrossRef.
  49. E. L. Fornero, A. L. Bonivardi and M. A. Baltanás, J. Catal., 2015, 330, 302–310 CrossRef.
  50. K. Li and J. G. Chen, ACS Catal., 2019, 9, 7840–7861 CrossRef.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta03275j
These authors contributed to this paper equally.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.