Abstract
Interface engineering is crucial in the design of supported metal catalysts, as it significantly influences the catalytic process, particularly in terms of selectivity. Herein, we discover the redispersion of Ni nanoparticles from SiO2 to molybdenum carbide (Mo2C) being induced by the strong interaction between metal and Mo2C. Parameters affect such migration are thoroughly investigated from carbon source, proximity between Ni and Mo2C to activation atmosphere and temperature. The established dual interface on Mo2C-Ni/SiO2 catalyst exhibits excellent catalytic performance for CO2 hydrogenation, readily shifting the selectivity from 91% CO on Ni/Mo2C to near 100% CH4 on Mo2C-Ni/SiO2. Density functional theory calculations further verify the interfacial synergy between Ni/Mo2C and Ni/SiO2 sites with low barrier for CO2 activation and a subsequent successive hydrogenation of CO. These findings highlight the important role of strong metal-support interaction (SMSI) induced metal redispersion for the rational fabrication of dual interfaces, leading to a highly active catalyst for target products.
Subject terms: Heterogeneous catalysis, Chemical engineering, Catalytic mechanisms
Engineering catalyst interfaces plays a key role in designing supported metal systems. Here, the authors report that Ni nanoparticles redisperse from SiO₂ to molybdenum carbide, forming dual interfaces that enhance tandem CO₂ hydrogenation.
Introduction
Supported metal catalysts are widely used in many industrial processes for the sustainable production of fuels and chemicals1. Typically, the interaction between metal and support directly determines the structure and properties of the resulting interfaces, thereby exerting a substantial influence on catalytic processes, with selectivity particularly being affected2–4. Therefore, much attention has been given to the precise tuning of the metal-support interface to attain desired products.
Distinct from the classical strong metal-support interaction (SMSI) effect, which typically characterized by reducible oxide encapsulation of metal nanoparticles, recent advances have expanded the scope of SMSI behaviors. Among these, metal redispersion represents a prominent SMSI-induced phenomenon under activation or reaction conditions5,6. Note that the SMSI as the driving force, is able to fragment large metal cluster into high density assemblies by various approaches, such as oxidative fragmentation7,8 and atomic trapping9,10. As a result, the interfacial contact between metal and support undergoes change, influencing the activity and selectivity of synthesized catalysts11. For example, a pioneering work performed by Jones et al.9 has revealed the restructuring of Pt on CeO2 surface during oxidative activation, and emphasized the importance of Pt-CeO2 interaction to incur the migration and redispersion of Pt via gas-solid route. Eventually, the movable Pt atoms could be effectively trapped by CeO2 support to suppress sintering, resulting in an improved activity in CO oxidation as well as prolonged durability. In another study, Paredis et al.12 observed the NO-induced redispersion of Pd over ZrO2 at the low temperature range (≤120 °C)) of NO reduction process, leading to a shifted selectivity toward N2O. These studies indicate that both the supporting material and redox environment play a pivotal role in the evolution of metal structure, thus influencing the performance of the reaction.
Transition-metal carbides (TMCs), molybdenum carbides (MoxC) in particular, have emerged as promising substrates for stabilizing metal nanoparticles (NPs), typically forming two characteristic configurations: epitaxial clusters with few atomic layer thickness13–15 or isolated single atoms16–21. Distinct from conventional metal oxides, the TMCs exhibit stronger interfacial chemical bonding with loaded metals under high-temperature reductive conditions22–27, a unique attribute that drives structural reorganization through SMSI. Notably, this SMSI-driven redispersion phenomenon enables metal migration between different substrates. For example, our previous study demonstrated the redispersion of Pt NPs from carbon nanotubes (CNTs) to α-MoC1-x substrate under reductive treatment28. Subsequent work revealed a significant decrease in Ni particle size along with the altered electronic properties of relocated Ni species on the Mo2C-Ni/Al2O3 catalyst, achieving 38% activity enhancement in dry reforming of methane29,30. Similarly, Zhou et al. demonstrated Cu migration from SiO2 to Mo2CTx, forming atomically dispersed Cu-Mo2CTx interfaces that boost methanol synthesis rates31. This analogous behavior also extends to nitride systems, in which the loaded Ni NPs undergo structural evolution into undercoordinated layer-like structure on γ-Mo2N substrate32. Despite these advances, three critical knowledge gaps remain: (i) the atomic mechanisms governing the SMSI-driven metal redispersion between two distinct substrates are elusive; (ii) current systems face intrinsic limitations in reconciling high metal loading with atomic-level dispersion; (iii) the catalytic implications of dual interfacial systems are underexplored. To address these challenges, further investigation is required. Therefore, as motivated by this intriguing behavior, there is a call for studies on dual interfacial effect to gain precise control over the underlying mechanism of metal migration across supports as well as the influence of major factors, ultimately enabling the tailored catalytic systems to obtain target products in cascade reactions.
Herein, we report how to optimize the selectivity/activity in CO2 hydrogenation by leveraging the SMSI effects to engineer the degree of metal dispersion as well as the amounts of metal-oxide/carbide interfaces. The Mo2C-Ni/SiO2 catalysts have been fabricated and taken as model systems for catalyzing methanation in CO2 hydrogenation. Effects of various factors (carbon source, mixing method, activation temperature and atmosphere) that influenced the redispersion of Ni NPs from oxide to carbide surface are investigated in attempt to understand the driving forces behind the migration as well as discriminate the key factors for such redispersion. The process of Ni redispersion from SiO2 to Mo2C substrate was confirmed through direct visualization of structural evolution by high-resolution transmission electron microscopy (HRTEM) and electronic state characterization by X-ray photoelectron spectroscopy (XPS). This combined evidence demonstrates the redispersion of Ni NPs from oxide to carbide, leading to the generation of new Ni/Mo2C interfaces. The synergy between the dual interface of Ni-Mo2C and Ni-SiO2 results in a doubled increase of CO2 conversion and significantly shifts the product selectivity, from 46% CO over Ni/SiO2 and 91% CO over Ni/Mo2C to 97% CH4 over the Mo2C-10Ni/SiO2 catalyst. The CH4 formation rate on the optimized Mo2C-10Ni/SiO2 catalyst (127 μmol gcat−1 h−1 ) is about 4.2 and 18.1 times higher than that of the 10Ni/SiO2 (30 μmol gcat−1 h−1 ) and Ni/Mo2C (7 μmol gcat−1 h−1 ) catalysts, respectively. After detailed characterizations and density functional theory (DFT) calculations, we discover that the observed changes in activity and selectivity originate from the tandem catalysis occurring at the established dual interfaces of Ni-Mo2C and Ni-SiO2, in which the activation of CO2 initially occurs on the Ni/Mo2C sites, resulting in the formation of CO* and subsequently, these CO* intermediates were successively hydrogenated on Ni/SiO2 sites to produce CH4. Findings from this study greatly broaden the applicability of the SMSI effects and provide insights into a promising approach to enhance catalysis by constructing dual interfaces.
Results
Evidence and driving force for Ni migration from SiO2 to Mo2C
To evidence and elucidate the intrinsic driving force for Ni redispersion, a systematic study was performed by examining the influence of various factors (carbon source, mixing method, activation temperature, and atmosphere) involved in the preparation of Mo2C-Ni/SiO2 catalysts. Firstly, the activated carbon (C) and Mo2C were individually chosen and mixed mechanically with the 10Ni/SiO2 in a mass ratio of 1–3 before undergoing the reduction treatment with a gas mixture of 15%CH4/H2 at 590 °C. The transmission electron microscopy (TEM) analyses were employed to determine the cluster size of Ni. It was found that a small portion of Ni aggregated unevenly on SiO2 over the Ni/SiO2 reference, with an average particle size of 19.5 ± 0.4 nm (Figs. 1a and S1). In contrast, the Ni-Mo2C/SiO2 catalyst exhibited well-distributed Ni-enriched dark dots on SiO2, characterized by a decreased particle size of 13.8 ± 0.2 nm (Figs. 1b and S2). While the estimated Ni particle size is about 18.1 ± 0.2 nm on the C-Ni/SiO2 catalyst (Fig. S3), indicating a negligible impact on Ni distribution compared to the Ni/SiO2 reference. Therefore, we can deduce that carbon (C) does not possess similar ability as Mo2C does, an indication of weak interaction between Ni and C. Figure 1c shows the Rietveld-refined powder X-ray diffraction (PXRD) profiles of the prepared Ni-based catalysts. Notably, the Mo2C-Ni/SiO2 catalyst exhibited a broad diffraction peak with reduced intensity on Ni (111) reflection plane compared to the Ni/SiO2 reference. The crystallite size of Ni decreased from 21.6 to 17.3 nm by calculating the reflection of Ni (111) plane (Fig. S4). Meanwhile, the estimated crystallite size of Ni obtained from CO pulse measurement showed a decrease upon mixing with Mo2C (Table S1). The X-ray photoelectron spectroscopy (XPS) spectra of Mo 3d5/2 (binding energy, B.E. = 227.7 eV) and Ni 2p3/2 (B.E. = 855.1 eV) clearly evidenced the formation of Mo-C-Ni motifs (Fig. 1d–f)29,33. The binding energy (B.E.) of Mo 3d5/2 being characteristic of Mo2C negatively shifted from 228.2 to 227.7 eV over the Mo2C-Ni/SiO2 catalyst, and correspondingly the B.E. of Ni 2p3/2 being ascribed to metallic Ni0 showed a positive shift from 852.7 to 853.0 eV. The newly emerged peak at 855.1 eV in the Ni 2p3/2 spectra, except for Ni0 and Ni2+ (856.3 eV, NiO) was assigned to Ni-Cx motif as confirmed in our previous studies29,30. Notably, the Ni2+ content in the Mo2C-Ni/SiO2 catalyst (22%) exhibited a marked decrease compared to that in the pristine Ni/SiO2 catalyst (47%). This reduction can be rationalized by the interfacial synergy between Mo2C and Ni as induced by the migration of Ni from SiO2 to Mo2C, which diminishes the anchoring effect of SiO2 support on Ni species. Consequently, the weakened metal-support interaction promotes the reduction of Ni2+ during catalyst activation process. This phenomenon aligns with our prior observation on the Mo2C-Ni/Al2O3 system29. Meanwhile, the low-valence Moδ+ (0 < δ < 2) species at B.E. of 227.7 eV were detected on the Mo2C-Ni/SiO2 catalyst, an indication of weakened Mo-C bonds, as a result of C being linked with Ni and the creation of Ni-Cx motifs. The assignment of 227.7 eV peak to Moδ+ aligns with our established methodology in an analogous system29. Such valence change of Ni and Mo in the synthesized composite catalyst indicates the charge transfers from Ni to Mo2C. The quantitative peak-deconvoluted data revealed that about 36% of Ni remains at electron-deficient states (Niσ+) after CH4/H2 activation on Mo2C-Ni/SiO2 catalyst (Table S2). The relative amount of Moδ+ (0 < δ < 2) increases to 44% after obtaining electrons. Thereafter, the relative amounts of Niσ+ and Moδ+ (0 < σ < 2, 0 < δ < 2) have been defined as descriptors to express the amounts of generated Ni-Mo2C interfacial sites. Different from Mo2C-Ni/SiO2, the C-Ni/SiO2 catalyst showed the formation of nickel carbides after CH4/H2 activation with the B.E.s of Ni 2p at 854.4 eV and C 1 s of 280.9 eV due to the strong Ni-C bonding (Fig. 1f).
Fig. 1. Structure and electronic states characterization of the as-prepared χ-Ni/SiO2 catalysts.
a TEM image of Ni/SiO2 catalyst. b TEM image of Mo2C-Ni/SiO2 catalyst. c PXRD patterns of Ni/SiO2, Mo2C-Ni/SiO2, and C-Ni/SiO2 catalysts with enlarged images marked in yellow. d–f detailed quasi in-situ XPS deconvolution of Mo 3 d, Ni 2p, and C 1 s spectra. g–i XPS semi-quantitative analysis of Mo2C-Ni/SiO2 prepared by different mixing methods, reduction temperatures, and gas atmospheres.
The effects of mechanical strength of different mixing methods on the preparation of Mo2C-Ni/SiO2 catalysts were examined. Noticeably, the crystallite size of Ni obtained by ball milling (BM) and motor milling (MM) prepared Mo2C-Ni/SiO2 catalysts exhibits reduced particle size compared to their granular stacking (GS) counterparts across the (111) plane (BM: 17.3 nm; MM: 18.7 nm; GS: 20.6 nm), underscoring the critical role of mechanical strength on Ni dispersion (Fig. S5). Importantly, the TEM-derived particle size shows a reasonable agreement with the XRD-derived crystallite dimensions (Fig. S6, Table S3), which validates the reliability of our multi-scale characterization approach. The difference of Ni particle size between BM and MM-prepared samples may arise from the varied distance between Ni/SiO2 and Mo2C. In order to rule out the possible physical structure change caused by mixing process, a control experiment was conducted by ball-milling 10Ni/SiO2 catalyst. It is worth noting that the crystallite size of Ni remained on the BM-treated 10Ni/SiO2 catalyst compared to the as-synthesized one (Fig. S7). Also, the XPS results showed that the relative contents of formed Ni-C species and the ratio of Niσ+/Ni0 (0 < σ < 2) and Moδ+/Mo2+ (0 < δ < 2) of BM-prepared Mo2C-Ni/SiO2 are higher than that of the MM-prepared Mo2C-Ni/SiO2. Such findings imply that the enhanced mechanical strength promotes the proximity between Ni and Mo2C sites, thus facilitating the migration of Ni and leading to an increased formation of Ni-Mo2C interfacial sites. A linear correlation of Moδ+/Mo2+ or Niσ+/Ni0 with Ni crystallite size was established as shown in Fig. S8. Thus, we can infer that the proximity between Ni and Mo2C is able to be adjusted from micro- to nano-scale based on the mechanical strength provided by different mixing methods, which in turn influences the dispersion of Ni NPs.
Next, the effects of gas atmosphere and reduction temperature as key factors were investigated. Both the TEM measurement and XRD profiles of Mo2C-Ni/SiO2 catalysts showed the decrease of Ni particle size under H2 or CH4/H2 atmosphere, which demonstrates the redispersion of Ni under reducing gases (Figs. S9 and S10, Table S4). Meanwhile, the quasi in-situ XPS spectra showed a comparable distribution of valence states for the Mo2C-Ni/SiO2 catalyst treated with H2 and CH4/H2, which is significantly distinct from the N2-treated one. (Figs. 1f and S11). Also, the TEM analyses revealed a reduction in Ni particle size from 19.5 ± 0.4 nm for the pristine Ni/SiO2 catalyst to 14.7 ± 0.2 nm following mixing with Mo2C and CH4/H2 treatment at 500 °C. Elevating the activation temperature to 590 °C resulted in a further decrease to approximately 13.8 ± 0.2 nm (Fig. S12). However, the crystallite size of Ni significantly increased to 19.9 ± 0.3 nm when reduced at 650 °C (Fig. S13, Table S5). This result suggests the effectiveness of Mo2C on keeping Ni NPs from growing at an appropriate temperature range. However, the severe coalescence of Ni NPs on SiO2 carrier was not able to be compensated by Ni redispersion above 590 °C. The XPS spectra (Fig. 1g, Table S2) revealed a temperature-dependent evolution of metal species. As the activation temperature increased from 500 to 590 °C, the Moδ+/Mo2+ ratio progressively increased from 1.59 to 1.67, reaching a plateau above 590 °C. Concurrently, the Niσ+/ Ni0 ratio showed a parallel enhancement from 0.81 to 0.85 within the same temperature window, indicative of Mo2C-mediated interfacial interactions facilitating Ni redispersion. However, a decline in the Niσ+/Ni0 ratio was observed between 590 and 650 °C, because the thermodynamic driving force for Ni redispersion is overtaken by Ni0 agglomeration on SiO2 support. This inversion suggests that while elevated temperatures initially promote the formation of Ni-Mo2C interfacial sites through the enhanced metal migration kinetics, beyond 590 °C the dominant mechanism shifts to Ni agglomeration on SiO2 surface, where the cohesive energy of metallic Ni clusters overcomes the dispersion forces generated by the Mo2C matrix34. Furthermore, the diffuse reflectance infrared Fourier-transform spectra (DRIFTS) were employed by using CO as a probe molecule to provide additional evidence on the electronic states of Ni species. The emerged Niσ+ adsorption features in 2120 ~ 2170 cm−1 displayed a distinct blueshift under identical thermal treatment conditions (Figs. S14 and S15)35. Overall, the above analysis suggests that the metallic Ni0 species become activated on the inert SiO2 surface when the activation temperature is beyond its Hüttig temperature (THüttig = 518 K). Also, part of the added Mo2C was slightly reduced because of the electron transfer from the relocated Ni species that reside on the surface of Mo2C36–38.
As illustrated in Fig. 2a, the Mo2C-Ni/SiO2 catalyst was prepared using a two-step synthesis method. Three distinct areas surrounding a Mo2C patch were randomly chosen and labeled as Regions I to III in Fig. 2b. Moreover, the high-magnification high-angle annular dark-field (HAADF-STEM) images with the simultaneous energy-dispersive X-ray spectroscopy (EDX) analyses, clearly revealed the spatial distribution of Ni and Mo species (Figs. 2c–e and S16–S18). Noted that the elemental mappings showed a pronounced distribution of Ni within Si- or Mo-enriched regions, as exemplified in Region (I). Specifically, Ni and Mo signals were well correlated in the bright region above the boundary line in Fig. 2c, indicating that Ni species were highly dispersed at high density on Mo2C surface. This observation is consistent with the EDX line-scanning results, in which the Ni signal increased in tandem with the Mo signal (Fig. 2f). In contrast, Ni species exhibited a more disordered distribution as small patches on the Si-enriched regions below the bright area. Fig. 2d, e further substantiated the distinct distribution of Ni on Mo2C and SiO2. Overall, this observation provides a compelling evidence of Ni redispersion from SiO2 substrate onto the Mo2C surface during the synthesis of Mo2C-Ni/SiO2 catalyst. To further verify the presence of Ni on Mo2C, an acid treatment (3 vol% HF) was employed to selectively dissolve the Ni/SiO2 component from the Mo2C-10Ni/SiO2 catalyst, thus confirming the formation of Ni-Mo2C interfacial sites (Fig. S19). Additionally, EDX line-scanning analysis across the Mo2C-Ni/SiO2 interface boundary tacked the progressive evolution of Ni distribution with increasing treatment temperature. The resulting line profiles corroborated the XPS data, demonstrating a rise in the Ni/Mo ratio from 0.11 at 500 °C to 0.21 at 590 °C, followed by stabilizing at temperature exceeding 590 °C (Figs. S20–S23).
Fig. 2. Electron microscopic characterizations of the as-prepared Mo2C-Ni/SiO2 catalyst (BM, 590 °C, CH4/H2).
a Schematic illustration of Ni NPs redispersion during CH4/H2 treatment induced by Mo2C. b–e STEM-HAADF images and the corresponding X-ray energy-dispersive spectroscopy (EDS) elemental maps of Ni, Mo, and Si (b: scale bar: 50 nm; c: scale bar: 10 nm; d, e: 20 nm). f EDX-line scaning of Mo2C-Ni/SiO2 catalyst.
Electronic interaction upon metal redispersion
The X-ray adsorption fine structure (XAFS) studies were performed to elucidate the electronic states of Ni species. All samples were properly treated in a glove box without exposure to air. Herein, the Mo2C-1Ni/SiO2 composite catalysts and 1Ni/SiO2 reference with 1 wt% of Ni were evaluated, as it is more conductive to obtain precise structural and chemical information from XAFS analysis. The Ni/SiO2 reference exhibits both characteristic metallic Ni0 and oxidized Ni2+, with electronic configurations aligning with the XPS results (Fig. S24). Also, the comparative analysis reveals that the low-temperature activated Mo2C-Ni/SiO2 catalyst (denoted as Mo2C-1Ni/SiO2-LT, LT = 400 °C) maintains electronic signatures nearly identical to that of the Ni/SiO2 reference, which confirmed the absence of interfacial electronic coupling without adequate thermal activation. The Ni K-edge (8333 eV) X-ray absorption near-edge structure (XANES) spectra (Fig. 3a) of high-temperature activated Mo2C-1Ni/SiO2 catalyst (denoted as Mo2C-1Ni/SiO2-HT thereafter, HT = 590 °C) showed a positive shift of Ni K absorption edge compared to the low-temperature treated Mo2C-1Ni/SiO2 catalyst. Also, the intensity of “white-line” (8350.5 eV for Ni K edge) increased with increased activation temperature. Both phenomena indicate the enhanced charge transfer resulting from the formed Ni/Mo2C interface, which is similar to the reported Co/γ-Mo2N and Pt1/FeOx catalysts39,40. Since most of the Mo ions still remain as carbide in Mo2C-1Ni/SiO2 catalysts, there is no obvious change on the Mo XANES (20000 eV) spectra (Fig. 3b). From the above analysis, we can deduce that the strong interaction between Ni and Mo2C is able to induce the electronic perturbation of Ni species, leading to the formation of charge-redistributed Ni-Mo2C interfaces, eventually the re-dispersed Ni could stay in cationic state. Similar phenomenon is also observed on the redispersion process of Pt or Rh NPs on CeO2 substrate under oxidative atmosphere9,41.
Fig. 3. Structure and electronic states characterization of Mo2C-Ni/SiO2 catalyst.
a Normalized Ni K-edge (8333 eV) XANES spectra of Mo2C-1Ni/SiO2 catalysts with different reduction temperatures. (Both the “white line” and “edge” are marked by a dashed line arrow; Ni foil and NiO were used as references) b Normalized Mo K-edge (20,000 eV) XANES spectra of Mo2C-1Ni/SiO2 catalysts. (Mo foil and Mo2C were used as references). c Fourier transform Ni K-edge EXAFS spectra of Mo2C-1Ni/SiO2 catalysts. d Fourier transform Mo K-edge EXAFS spectra of Mo2C-Ni/SiO2 catalysts. e, f Ni K-edge wavelet transformation of Mo2C-1Ni/SiO2-LT and Mo2C-1Ni/SiO2-HT catalysts, respectively. (The insert of Fig. 2f is the contour plot of charge density difference upon the formation of Ni1@Mo2C sites; the charge accumulation and depletion regions are colored by yellow and blue, respectively).
Moreover, the Ni K-edge extended X-ray absorption fine structure (EXAFS) was employed to study the local chemical environment of Ni species. The fitting results (Figs. 3c and S25, Table S6) revealed the presence of Ni-C, Ni-Mo, and Ni-Ni coordination on the Mo2C-1Ni/SiO2-HT catalyst. The mean bond length of Ni-C is about 2.07 Å, while the bond length of Ni-Mo is around 2.91 Å. According to a previous study20, we can deduce that the Mo atoms are located beneath the subsurface layer of C. Notably, the average coordination number of Ni-Ni (CNNi-Ni) decreased from 4.3 to 1.9 as the reduction temperature increased over the prepared Mo2C-1Ni/SiO2 catalyst. This result again confirmed the redispersion of Ni on Mo2C, which can occur at HT, and the anchored Ni atoms bond to Mo2C surface via C as bridges. In addition, Mo K-edge (20,000 eV) spectra were analyzed (Figs. 3d and S26, Table S7). Obviously, the CNMo-C-Mo decreased with increased reduction temperature from 9.5 to 8.7, perhaps due to the formation of Ni-C-Mo bonds, thus reducing the intensity of the second-shell scattering of Mo2C. This result is in accordance with the Ni K-edge spectra. The EXAFS fittings were further validated by a wavelet transformation (WT) of the EXAFS spectra. The WT contour maps of the Mo2C-1Ni/SiO2-LT and Mo2C-1Ni/SiO2-HT catalysts are distinct, as shown in Fig. 3e, f, respectively. The component in the R range near 1.6 Å is attributed to the coordination between Ni and C/O, whereas the R at 2.1 Å is contributed by the Ni-Ni coordination based on the similar k-space intensity distribution. Interestingly, the R range appeared at ca. 2.6 Å belongs to the Ni-Mo bond, indicating the significant role of activation temperature on the extent of Ni NPs redispersion.
In addition, the ab-initio thermodynamic calculations proved that the Ni NPs prefer to be atomically dispersed over the Mo2C(101) plane with a relatively large formation energy of −5.41 eV (Fig. S27a). The population analysis verified the charge transfer between Ni and Mo2C (Ni: +0.18|e | ) and confirmed the energetically favorable connection between Ni and Mo2C via Ni-C-Mo bonds. (Fig. S27b–d), Figs. S28–S30, Tables S8–S10).
Catalytic performance in CO2 hydrogenation
The catalytic performance of Mo2C-10Ni/SiO2 catalyst and several references was evaluated in CO2 hydrogenation due to the structure-sensitive nature of this reaction42–45. As shown in Fig. 4a, b, the 10Ni/SiO2 catalyst exhibited a CO2 conversion of 20.9% with CH4 selectivity of 54% at 350 °C with a WHSV of 30,000 mL g−1 h−1. In addition, both the Mo2C and Ni/Mo2C references showed an enhanced CO2 conversion due to their unique atomic arrangement and platinum-like behaviors, endowing an effectiveness for CO2 adsorption and activation. Note that the Mo2C-based catalysts presented an ultrahigh CO selectivity up to 91%, which is consistent with Zhang et al.17 observation. However, the limited surface areas of β-Mo2C could impede the maximum Ni loading. How to resolve the intrinsic trade-off of metal loading and dispersion becomes an issue in catalyst design. To address this, we designed a Mo2C-Ni/SiO2 composite catalyst with Ni loading of 10% by taking advantage of the redispersion of Ni from SiO2 to Mo2C via the SMSI between Mo2C and Ni. Impressively, the CO2 conversion of the dual interfacial Mo2C-10Ni/SiO2 catalyst is the highest, more than two times that of the 10Ni/SiO2 catalyst. Moreover, the selectivity shifts from CO over Mo2C (~92%) and 10Ni/SiO2 (~45%) towards CH4 (~97%) over the Mo2C-10Ni/SiO2 catalyst at the same CO2 conversion level (Fig. S31a). In contrast to Mo2C, there is no improvement on the activity while adding C as additive (Fig. S32). Therefore, we propose that the enhanced CH4 yield may result from the synergistic interplay between Ni-Mo2C and Ni-SiO2 interfaces that single interfacial system cannot be obtained. To further confirm this, we prepared another reference sample by physically mixing 10Ni/SiO2 with 1Ni/Mo2C without thermal pretreatment in CH4/H2 (denoted as 10Ni/SiO2 + 1Ni/Mo2C). It was found that the two interfaces that involved in the 10Ni/SiO2 + 1Ni/Mo2C reference (CO2 conv. = 45.1%; CH4 selec. = 95.5%) work synergistically in CO2 hydrogenation, resulting in a comparable performance to that of the Mo2C-10Ni/SiO2 catalyst (CO2 conv. = 48.6%; CH4 selec. = 97.1%). Therefore, we can conclude that Ni redispersion induced by the SMSI between Mo2C and Ni results in the formation of Ni-Mo2C interfaces, establishing highly-dispersed catalysts at high metal loading with dual interfacial sites for CO2 hydrogenation, which breaks the limitation of conventional preparation method.
Fig. 4. Catalytic performance of Mo2C-Ni/SiO2 catalysts in CO2 hydrogenation.
a CO2 conversion and CH4 STY of Mo2C-10Ni/SiO2 catalyst with references at 350 °C, 0.1 MPa, and 30,000 mL g−1 h−1 with CO2:H2 ratio of 1:4. b Product (CO or CH4) selectivities. c Effect of mixing methods in CO2 hydrogenation. (DL-double layer stacking; GS-granular stacking, MM motor milling, BM ball milling. error bar represents standard deviation; the error bars in a, b represent the standard deviation from three or more independent experiments). d The corresponding product selectivities of various mixing methods. e Calculated TOFs of CH4 formation over 10Ni/SiO2 and Mo2C-10Ni/SiO2 catalysts (300 °C, 0.1 MPa, 80,000 mL g−1 h−1 ). f The calculated activation energy (Ea) for CH4 formation of 10Ni/SiO2 and Mo2C-10Ni/SiO2 catalysts. (275–350 °C, 0.1 MPa, and 80,000 mL g−1 h−1 ).
Followed by this, the effects of the integration manner were investigated (Figs. 4c, d and S31b). It was found that both the CO2 conversion and CH4 selectivity increase in the order of double layer (DL-conv. 25.5%; CH4:65.4%) <granule stacking (GS-conv. 27.7%; CH4:72.2%) <motor milling (MM-conv. 38.2%; CH4:89.4%) <ball milling (BM-conv. 48.6%; CH4:97.1%), which is consistent with the increase in Ni-Mo2C interfacial sites. As a result, a higher degree of redispersion resulting from distinct mixing methods was able to generate more Ni-Mo2C interface, thus enhancing the formation of CH4. Aside from the mixing method, other factors were investigated as well, including the mass ratio of Ni/SiO2 to Mo2C, activation temperature, and atmosphere (see details in Figs. S33 and 34). Therefore, we can reasonably deduce that the relative ratio between Ni-Mo2C and Ni-SiO2 and the activation process significantly impact the activity and selectivity of CO2 hydrogenation by the SMSI-induced Ni redispersion. As depicted in Fig. S34, the obtained molar ratio of Ni/Mo was about 0.5, indicating that the redispersion degree of Ni was closely related to the amount of Mo2C. Also, the CO2 conversion exhibited a gradual increase as the activation temperature increased up to 590 °C and slightly dropped thereafter (Fig. S34). Only a good match between CO2 activation and COx hydrogenation is able to enhance the CH4 formation rate.
To ensure the measurements were maintained in a kinetic-relevant regime, the CO2 conversion was limited to 15% with a relatively high space velocity of 80,000 mL g−1 h−1 at various temperatures. The calculated approach to equilibrium (η) is about 0.16, which aligns with the recommended threshold (η < 0.2) for kinetic control and suggests that the experimental condition applied is far from equilibrium46. All rate measurements were thermodynamically corrected, which confirmed the kinetic control throughout the experimental range (Fig. S35, Tables S11–S14). As displayed in Fig. 4e and Table S15, the calculated turnover frequency (TOF; normalized by the number of CO chemisorption sites) of CH4 formation of the Mo2C-10Ni/SiO2 catalyst (7.1E-04 s−1 ) is about 7.8x’s higher than that of the 10Ni/SiO2 catalyst (9.0E-05 s−1 ), indicating the improved intrinsic activity of Ni sites due to their altered electronic properties. Taking the observed Ni redispersion from SiO2 to Mo2C and the established electron transfer into account, we propose that the increased TOF originates from synergistic effects combining better Ni exposure and optimized electronic structure. Moreover, the apparent activation energy (Ea) of Mo2C-10Ni/SiO2 and 10Ni/SiO2 catalyst for CO2 hydrogenation was determined to be 67.7 ± 1.1 and 85.2 ± 0.7 kJ mol−1 , respectively (Table S12). The obtained Ea over 10Ni/SiO2 catalyst is in good agreement with the literature reported value47–49. Importantly, the activation energy (Ea) of CH4 formation exhibited a substantial reduction from 93.8 ± 0.6 kJ mol−1 over Ni/SiO2 to 67.6 ± 1.1 kJ mol−1 for the Mo2C-Ni/SiO2 catalyst, demonstrating that Mo2C incorporation effectively lowers the energy barrier for methane generation (Fig. 4f, Table S16). In contrast, the Ea decreased for CO formation was less pronounced, shifting from 79.8 ± 0.7 to 72.4 ± 3.2 kJ mol−1 upon Mo2C addition (Fig. S36). This differential effect highlights the preferential enhancement of CH4 pathway kinetics by Mo2C-containing sites. The variation in the activation barrier of 10Ni/SiO2 and Mo2C-Ni/SiO2 catalysts implies the facile activation of CO2 on Mo2C-Ni/SiO2, which may result from the Ni redispersion-induced electronic and geometric structure change on active Ni sites.
Reaction mechanism exploration
As shown in Fig. 5a–h, we performed a two-step temperature-programmed surface reaction (TPSR) to explore the reaction mechanism of CO2 hydrogenation over the Ni/SiO2, Mo2C-Ni/SiO2, and Mo2C catalysts. Initially, a 2%CO2/He mixture (50 mL min−1) was introduced at 300 °C after activation. Noted that a small amount of CO(g) was immediately produced on the Ni/SiO2 catalyst (Fig. 5a). While no gaseous CO was detected over the Mo2C-Ni/SiO2 catalyst (Fig. 5d), either because of the strong interaction of CO* to the active sites or an alternative CO2 activation route. To verify the potential intermediates, we performed a temperature-programmed desorption (TPD) under He flow after CO2 dissociation (Fig. 5c, f). Notably, the attached intermediates gradually desorbed as gaseous CO with increased temperature on both catalysts. For Ni/SiO2 catalyst, a broad CO desorption peak was detected, which starts at around 400 °C; while only one CO desorption peak appeared with the onset temperature of 500 °C on the Mo2C-Ni/SiO2 catalyst, which implies that the interaction between CO* and Mo2C-Ni/SiO2 surface is stronger than that of the Ni/SiO2 catalyst. Also, the amount of desorbed CO on Mo2C-Ni/SiO2 is about two times that of the Ni/SiO2 catalyst, indicating that more CO* species are preserved on the former catalyst. In sharp contrast, the Mo2C alone exhibited a spontaneous CO2 dissociation, forming CO(g) as a major product, indicating the weak interaction between CO and the O* occupied Mo carbide surface (Fig. 5g). This result is similar to that of the Ni/Mo2C reference (Fig. S37). Followed by this, the system was switched to 5%H2/He (50 mL min−1 ) at the same temperature after purged with N2 (50 mL min−1 , 30 min) (Fig. 5b, e). Note that the amount of generated CH4 in the 2nd step of hydrogenation over Ni/SiO2 was less than that of Mo2C-Ni/SiO2, which is consistent with the TPD results as well as our experimental data. It suggests that the produced CH4 mainly originates from the hydrogenation of adsorbed CO. Similar to the Mo2C-Ni/SiO2 catalyst, the physically mixed 10Ni/SiO2 + 1Ni/Mo2C reference also exhibits a considerable formation of CH4 as a result of the two-step tandem process on the involved dual interface (Fig. S38). Meanwhile, the in-situ DRIFT experiments were carried out to probe surface intermediates by co-feeding CO2 and H2 mixture (1:4) over the Mo2C-Ni/SiO2 catalyst (Fig. S39). It was found that more CO* intermediates are detected at low temperature, as indicated by the bands located at 2141, 2075, 2030, and 2010 cm−1 , which are attributed to the adsorbed CO* on the Niσ+, Ni0, and Mo2C sites over the Mo2C-Ni/SiO2 catalyst, respectively. The emergence of CO* indicates the strong CO2 dissociation ability of the Mo2C-Ni/SiO2 catalyst, even at a relatively low temperature. This result is in agreement with the two-step MS experiment. Also, the intensity of CH4 signal at 3016 cm−1 increases with increased reaction temperature, most likely due to the successive hydrogenation of CO*.
Fig. 5. Insights into the reaction mechanism over Mo2C-Ni/SiO2 catalyst.
a–h Two-step MS investigation of Ni/SiO2 (a–c), Mo2C-Ni/SiO2 (d–f), and Mo2C (g, h) catalysts. CO2 dissociation at 300 °C (a, d, g), followed by H2 treatment (b, e, h) (Products were detected by online mass spectrometer) or a temperature-programed desorption under He flow (TPD: c, f). i Energy barrier (Ea) for CO2 and H2 dissociation on Ni4/SiO2, Ni1/Mo2C (101), and Mo2C models. j Adsorption energy of CO. k Reaction paths for CO2 hydrogenation over Mo2C-Ni/SiO2 catalyst.
To get more insights on the catalytic function of involved active sites, the Ni4/SiO2, Ni1/Mo2C(101), and Mo2C(101) models were built and used as references, representing Ni/SiO2, Ni/Mo2C, and Mo2C catalysts, respectively. The energy barrier of these three catalyst models for CO2 and H2 dissociation (Ediss,CO2 & Ediss,H2) was calculated and compared, as displayed in Fig. 5i. The Ediss,CO2 of Ni1/Mo2C(101) (0.46 eV) is lower than those of Ni4/SiO2 (1.83 eV) and Mo2C(101) (0.63 eV), indicating that the Ni/Mo2C catalyst is favorable for CO2 activation. The value of Ediss,H2 follows the trend of Mo2C (0.31 eV) > Ni1/Mo2C (0.17 eV) > Ni4/SiO2 (0.15 eV), indicating the outstanding performance of Ni4/SiO2 in H2 activation. We then investigated the important successive step by calculating CO adsorption energy (Eads,CO) over those three models (Fig. 5j). The results show that the Ni/SiO2 model binds CO much stronger (−2.30 eV) than Ni/Mo2C (−1.95 eV) and Mo2C(101) (−1.58 eV), which is consistent with the CO-TPD measurement. Apart from this, Fig. S40 displays the CO desorption behavior on Ni/SiO2, Mo2C, and Mo2C-Ni/SiO2 surface. Clearly, a new CO-desorption peak (ca. 242 °C) appeared on the Mo2C-Ni/SiO2 catalyst, indicating the formation of new active sites, which might originate from the Ni-Mo2C domains. Similar result was also observed by Ma et al.50 on a tandem Pt-Mo2C/C catalyst. Therefore, we can deduce that the electronic interfacial sites provided by Ni-Mo2C interface showed the potential of stabilizing undercoordinated Niδ+ species for the selective cleavage of C = O bond, while the structural interfacial sites of Ni-SiO2 maintained the high-density metallic ensembles for efficient H2 dissociation. The presence of Mo2C modulated the interaction between metal and support by preventing the over stabilization of intermediates on Ni/SiO2 sites. This bifunctional synergy, in which the Mo₂C dictates electronic states and SiO₂ controls nanostructure, resolves the persistent activity-selectivity-lifetime trade-off.
Extensive DFT-based calculations were performed to highlight the synergy between Ni/SiO2 and Ni/Mo2C dual interfaces in CO2 hydrogenation (Fig. 5k). Ni atoms are stabilized by two-fold coordinated C atoms on Mo2C (101) plane as shown on the insert of Fig. 3f. These partially charged and undercoordinated Ni atoms are more efficient than surface Mo as reaction sites for CO2 and H2 activation. The initial barrier for CO2 dissociation (CO2* + 4H* → CO* + O* + 4H*) on Ni/Mo2C interface is about 0.63 eV, which is comparable to that over bare Mo2C (101) surface (Fig. S41). Also, this step is 0.69 eV exothermic, forming stable Ni-carbonyl in-situ. Moreover, we notice that the hydrogenation of adsorbed O* can be regarded as the most sluggish step on the Ni/Mo2C (101) and Mo2C (101) surface with energy barrier of 1.18 and 1.49 eV (TS3 and TS5, Fig. 5k), respectively, but is vital to keep the reaction proceed. Originating from the Ni/Mo2C interface, the previously formed Ni-carbonyl sites could convert the adsorbed CO2 and H2 at an energy barrier of 0.46 and 0.17 eV (TS6 and TS4, Fig. 5k), respectively, which are significantly lower than those over Mo2C (101) plane and Ni/SiO2 interface. Furthermore, CO generated by CO2 dissociation will be shifted to a surface Mo atom that is not available for CO adsorption/desorption on Mo2C(101), where CO desorbs easily at a barrier of 0.85 eV (step 16-17) at reaction conditions. Compared to Mo2C(101), this value decreased about 1.0 eV. In contrast, we found that the rate determine step of CO2 hydrogenation over Ni4/SiO2 model is CO2 dissociation, which needs to overcome an energy barrier of 1.83 eV though the reaction is exthermic by 0.58 eV. These results indicate that CO2 can be activated easily on Ni/Mo2C interface to facilitate the breakage of C = O bonds as well as the CO desorption. Driven by the aforementioned strong CO adsorption ability, the released CO would migrate to Ni4/SiO2 interface for further hydrogenation, eventually forming CH4 as final product. Given the experimental and computational results, we can propose that the synthesized Mo2C-Ni/SiO2 catalyst constitutes a tandem catalytic system for CO2 hydrogenation as induced by the SMSI between Mo2C and Ni, in which the CO2 reduction could initially occur on the Ni/Mo2C interface while the subsequent hydrogenation of CO takes place on the Ni/SiO2 interface.
Discussion
In summary, we report the redispersion of Ni NPs from SiO2 substrate to Mo2C under reduction conditions. The established Ni-C-Mo bonds stabilize Ni on Mo2C surface in electron-deficient states as well as decrease the particle size of Ni nanoparticles. This results in the creation of Ni/Mo2C interface, which works synergistically with the original Ni/SiO2 interface to yield a dual interfacial system for CO2 hydrogenation. A detailed tandem catalytic mechanism, involving interfacial synergistic catalysis over the Mo2C-Ni/SiO2 catalyst, has been elucidated. Both the experimental studies and theoretical calculations substantiate that the Ni-Mo2C interfacial sites are catalytically active for CO2 activation, resulting in an increased formation of CO species, which are readily hydrogenated on Ni-SiO2 interface to produce CH4. This work opens up a new avenue for designing dual interface systems to enable a precise control on the reactivity and selectivity of catalytic process.
Methods
Catalyst synthesis
Preparation of Mo2C-Ni/SiO2 catalysts: The Ni-based catalysts (denoted as Ni/SiO2) were synthesized by incipient wetness impregnation (IWP) with Ni nominal loading in the range of 1–10 wt% on silica support. Molybdenum carbide (denoted as Mo2C) as the second component was synthesized by a temperature programmed reduction as follows: Firstly, the ammonium heptamolybdate ((NH4)6Mo7O24·4H2O) was heated to 500 °C in a muffle furnace and held for 4 h to prepare MoO3 precursor. After that, the obtained MoO3 was heated from room temperature (RT) to 300 °C in 20 vol% CH4/H2 mixture with a ramping rate of 5 °C/min, then increased from 300 to 700 °C at a ramping rate of 1 °C/min and held for 2 h at 700 °C. The resulting material was quickly quenched to room temperature and passivated by 1%O2/Ar mixture (20 mL/min) for 4 h. The investigated materials were further prepared by mixing Ni/SiO2 with Mo2C with a mass ratio of 5:1–1:3. Finally, the prepared mixture is subjected to a 15%CH4/H2 treatment at designated temperature (400–650 °C) for 2 h. The resulting samples were denoted as Mo2C-Ni/SiO2.
Preparation of Ni/Mo2C reference catalysts: Ni/Mo2C catalysts were prepared by co-precipitation and carburization for comparison. Prior to air exposure, the Ni/Mo2C catalyst was typically passivated in 1%O2/N2 flow for 2 h.
Catalyst characterization and CO2 hydrogenation evaluation
X-ray diffraction (XRD): Powder X-ray diffraction (XRD) data were collected using a Rigaku SmartLab 9 kW diffractometer with Cu Kα radiation (λ = 1.5406 Å) with a step size of 0.02° and counting time of 8 s per step in the range of 2θ = 10-80°. The Diffraction patterns were compared to powder diffraction files from the Joint Committee on Powder Diffraction Standards (JCPDS) database. The crystallite size was estimated from the broadening peak of (111) plan for Ni by Scherrer equation. All the catalysts were passivated before measured.
Quasi in-situ X-ray photoelectron spectroscopy (XPS): The information of surface structure of the as-prepared catalysts was collected by XPS of Axis Ultra Imaging Photoelectron Spectrometer (Kratos Analytical). The prepared Mo2C-Ni/SiO2 and their references were activated in a continuous flow fixed-bed quartz reactor and treated under the target feed gas. Treated catalysts were made into small tablets (6.0 mm diameter) and placed on the sample holder without exposure to air in a glovebox. The sample was then introduced into the ultra-high vacuum chamber for XPS measurements without air exposure. 284.6 eV was used as an internal binding energy standard of C 1 s to correct for surface charging shift. XPSPEAK41 was used for spectra deconvolution by a nonlinear least-squares method with a combination of Lorentzian (20%) and Gaussian (80%) peak shapes.
Electron microscopy characterization: The high-resolution transmission electron microscope (HR-TEM) images of the prepared catalysts were recorded on a (JEOL) JEM 2200FS electron microscope with an accelerating voltage of 200 kV. Samples were prepared by dispersing in ethanol and sonicated for 300 s. A drop of the suspension was placed on a 230-mesh copper grid coated with Formvar-carbon film and then dried in a vacuum chamber. The clusters/particles identified in the TEM images were measured and counted to yield a size distribution that was fitted to a Gaussian distribution (100–200 particles were counted for statistical analysis) to determine the average size of the cluster/particle. All image analysis was conducted with Digital Micrograph software.
CO-adsorbed diffuse reflectance infrared spectra (CO-DRIFTs) and in-situ DRIFT: DRIFT spectroscopy of CO adsorption was employed on a Thermo Nicolet iS50 spectrometer with a mercury cadmium telluride (MCT) detector cooled by liquid nitrogen to estimate the electronic state of Ni and Mo species. All catalysts were prepared by diluting with quartz sands at a mass ratio of 1:5 prior to DRIFTS measurements to minimize the intermolecular dipole interaction of adsorbed CO. All the tests were performed in Harrick Praying Mantis high temperature reaction chamber with ZnSe windows. A thermocouple mount that allowed direct measurement of sample surface temperature. Each recorded spectrum was averaged over 32 scans in the region of 4000-650 cm−1 with a nominal 32 cm−1 resolution. Prior to CO adsorption, the precursor was pre-treated by 15%CH4/H2 at 590 °C for 2 h, unless otherwise indicated. After that, the flow was switched to N2 until the temperature decreased to 17 °C at a flow rate of 40 ml (STC)/min to clean the catalyst surface and then a background spectrum was recorded. CO adsorption was conducted at 17 °C by introducing CO into the system for above 20 min to ensure the complete saturation of CO gas molecules on catalyst surface. Finally, N2 was introduced to flush the physiosorbed CO. Several spectra were collected by 30 s interval until the spectra were stabilized.
X-ray absorption fine structure (XAFS): Ni L3 edge (8333 eV) and Mo K-edge (20000 eV) XAFS were measured at BL14W beamline of Shanghai Synchrotron Radiation Facility (SSRF) in transmission/ fluorescence mode using Lytle detector to collect the data. Ni foil and NiO were used as standards for Ni K-edge spectra; Mo foil and MoC were used as standards for Mo K-edge spectra. All the samples were pre-activated in 10%H2/Ar at 350 °C for 1 hour and sealed in a chamber under argon protection in a glove box.
Two-step temperature programmed surface reaction (TPSR): The 1st step is CO2 dissociation—The activated samples were exposed to 2%CO2/He (50 mL min−1 ) at a proper temperature for about 10 min. The effluent gases were measured online using a Mass Spectrometer (MKS-Cirrus3, USA). After N2 puring (50 mL min−1 ), the samples were reduced in 5%H2/He (50 mL min−1 ) in the 2nd step hydrogenation. In the control experiment, the 2nd step hydrogenation was replaced by a temperature-programmed desorption (TPD) under He flow to observe the desorbed species.
CO temperature-programmed desorption (CO-TPD): Temperature programmed desorption of CO (CO-TPR) was carried out on Micromeritics AutoChem 2920. The sample (100 mg) was reduced at 400 °C by 5%H2/Ar flow (30 mL min−1 ) for 1 h to remove passivation layer. After cooling dow, the CO adsorption was performed at 40 °C for 1 h. The sample was then flushed with N2 (30 mL min−1 ) until the baseline is stable. The desorption was conducted from 40 to 650 °C with a ramping rate of 10 °C/min. Dosorbed species were monitored by mass spectrometer (MS).
CO Chemisorption: This technique was used to determine the number of active sites (#CO, μmol g−1 ) on a Quantachrome ChemBETPulsar analyzer. Samples were purged with He at 100 °C for 30 min to remove moisture, and then reduced by 5%H2/Ar flow at 350 °C for 1 h to remove the passivation layer. Afterwards, the CO uptake was measured using pulsed chemisorption at 35 °C by taking thermal conductivity detector (TCD) as detector. The turnover frequency (TOF), defined as the number of CO2 converted per surface site per second, was calculated based on the following equation:
| 1 |
where the FCO2, inlet represented the inlet flow rate of CO2 (mol s−1 ), M represents the molecular weight of CO2 (g mol−1 ).
To determine Ni particle size (assuming cubic geometry), CO chemisorption measurements were performed on a Micrometritics AutoChem III 2930. Approximately 100 mg of the sample was loaded into a U-shaped quartz tube. The sample was prepared in situ under different conditions. After cooling to 35 °C and purging with high-purity He, the TCD signal was allowed to stabilize for 30 min. The sample was then exposed to ten pulses of 10% CO/He. The quantity of CO adsorbed was detected by TCD by assuming a stoichiometric ratio of one CO molecule per surface Ni atom.
CO2 hydrogenation was evaluated in a packed bed reactor, in which 100 mg catalyst powder mixed with 300 mg pre-calcinated SiO2 were loaded into a quartz tube with an inner diameter of 6 mm. The reactant was composed of 12% CO2/48%H2/15%Ar in N2 with a total flow rate of 50 mL (STP)/min. Ar was used as the internal standard. The activity test was conducted at 0.1 MPa in the temperature range of 548 to 623 K with a GHSV of 30,000 mL g−1 h−1 . The products were analyzed online using an Agilent GC7890B equipped thermal conductivity (TCD) and flame ionization detector (FID), respectively. For the Mo2C-Ni/SiO2 samples, the standard pre-treatment was conducted at 590 °C for 2 h in 15%CH4/H2 with a flow rate of 50 mL (STP)/min; whereas, H2 was only used for Ni/SiO2 references. In the quantification, the conversion of CO2 is defined as:
| 2 |
The observed rate (robserved) is corrected using the Temkin-Pyzhev formalism as follows:
| 3 |
The carbon balance is defined as:
| 4 |
In all tests, the carbon balance of the system is above 94%.
Computational details
The spin-polarized first-principle calculations are performed with Vienna Ab Initio Simulation Package (VASP 5.4.4). Perdew-Burke-Ernzerhof (PBE) functional was used to solve the Kohn-Sham equations within periodic boundary conditions. The electron-nucleus interactions are described using PAW pseudopotentials. A 4 × 4 × 4 Monkhorst-Pack k-point grid was used for bulk β-Mo2C and Ni, and a 2 × 2 × 1 k-point grid was used for the layered SiO2 structure derived from a recently reported silicate. Γ-centered grids with equivalent k-point density were then used in calculations of clean and Ni-deposited surface structures on β-Mo2C and SiO2. A planewave cutoff of 400 eV was used throughout this work. Second-order Methfessel-Paxton smearing with a width of 0.10 eV was used to handle the partial occupancies of the bands. The convergence criterion for energy is 1 × 10−5 eV and that for ionic relaxation is 0.01 eV/Å. With the above setup, the calculated lattice parameters a, b and c of bulk β-Mo2C are 6.06, 6.07 and 4.07 Å, respectively, while that for bulk face-center-cubic Ni was 3.51 Å. The layered SiO2 was found plausible in an orthorhombic lattice of 10.94 × 10.30 Å2. After test calculations, (101) surface slab with thickness of 4 Mo layer with mixed C and Mo termination was selected to mimic (101) surface of β-Mo2C, as the adsorption energy of a single Ni atom on (1 × 1) slabs in the configuration bridging 2 nearest neighboring surface C already converges within 0.03 eV. The stability of deposited Nin structures was investigated on a (2 × 2) supercell of (101) β-Mo2C and a (1 × 1) orthorhombic slab of SiO2. The freestanding Nin clusters were investigated in a 15.00 × 15.10 × 15.20 Å3 orthorhombic cell. The relative stabilities of Nin structures () were determined with the normalized formation free energy, calculated as:
| 5 |
where , , and are the free energy of the Nin/support composite, the number of Ni atoms in the Nin/support composite, electronic energy of a single Ni atom and the free energy of the support, respectively. First-principles thermodynamics scheme was adapted to calculate free energy of Nin/support composites and the support at finite T. Free energy is approximated from Helmholtz free energy by ignoring pV term, and was calculated as:
| 6 |
Where and are vibrational energy and entropy, respectively. was calculated with partition functions of structures in the temperature range of interest. For reference, the formation energies of these Nin structures () were also calculated as:
| 7 |
where , , and are the total energy of the Nin/support composite, the number of Ni atoms in the Nin/support composite, electronic energy of a single Ni atom and the total energy of the support, respectively. With the above setup and definition, the calculated is −4.79 eV.
The pathways for CO2 hydrogenation over Mo2C(101), Ni1/Mo2C(101) and Ni4/SiO2 were investigated with extensive spin-polarized first-principles-based calculations using Perdew-Burke-Ernzerhof (PBE) functional within the formalism of generalized gradient approximation (GGA) as implemented in the DMol3 package. DFT semi-core pseudopotentials (DSPP) and the double numerical plus polarization (DNP) basis sets were used to describe the core and valence electronic states. Test calculations on bulk and surface structures of Mo2C, Mo2C(101), Ni/Mo2C(101) and Ni4/SiO2, as well as energetics and structural of CO, CO2 and H2 adsorption over these model structures yield essentially the same results as those obtained with VASP.
Supplementary information
Acknowledgements
C.S. thanks the National Key R&D Program of China (No. 2023YFA1506602), the National Natural Science Foundation of China (Nos. 21932002 and 22276023), the Fundamental Research Funds for the Central Universities (DUT22LAB602, DUT24RC(3)061), and Liaoning Binhai Laboratory Project (LBLF-202306). H.W. acknowledges the National Natural Science Foundation of China (22202028), the Fellowship of China Postdoctoral Science Foundation (No. 2022M720638), and the Talent Introduction Program of Post-doctoral International Exchange Program (Grant no. YJ20210238). The authors thank the beamline BL13SSW (31124.02.SSRF.BL13SSW) in the Shanghai Synchrotron Radiation Facility for providing beamtime for XAS analyses.
Author contributions
H.W.: conceptualization, investigation, methodology, formal analysis, data curation, writing—original draft, software, writing—review & editing, visualization, funding acquisition. X.Q.: XAS analysis, methodology, formal analysis, data curation, software, writing—review and editing. Z.G.: TEM operation and analysis, methodology, formal analysis, data curation, software, writing—review and editing. Y.D., S.L. (Shenghua Liu), M.P., S.L. (Shida Liu): data curation, writing—review and editing. S.H.: conceptualization, Supervision. X.L.: conceptualization, project administration. X.G., D.M., and C.S.: conceptualization, supervision, funding acquisition, project administration.
Peer review
Peer review information
Nature Communications thanks Tomás Vergara, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Data availability
The data generated in this study are available within the paper, its supplementary information files, and the repository. Source data are provided with this paper (Figshare 10.6084/m9.figshare.29924090). Data are available from the corresponding authors upon request.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
These authors contributed equally: Haiyan Wang, Xuetao Qin, Zirui Gao.
Contributor Information
Xin Liu, Email: xliu@dlut.edu.cn.
Xinwen Guo, Email: guoxw@dlut.edu.cn.
Ding Ma, Email: dma@pku.edu.cn.
Chuan Shi, Email: chuanshi@dlut.edu.cn.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-025-64030-9.
References
- 1.Schlögl, R. Heterogeneous catalysis. Angew. Chem. Int. Ed.54, 3465–3520 (2015). [DOI] [PubMed] [Google Scholar]
- 2.Pan, C.-J. et al. Tuning/exploiting strong metal-support interaction (SMSI) in heterogeneous catalysis. J. Taiwan Inst. Chem. Eng.74, 154–186 (2017). [Google Scholar]
- 3.Campbell, C. T. & Mao, Z. Chemical potential of metal atoms in supported nanoparticles: dependence upon particle size and support. ACS Catal.7, 8460–8466 (2017). [Google Scholar]
- 4.James, T. E., Hemmingson, S. L. & Campbell, C. T. Energy of supported metal catalysts: From single atoms to large metal nanoparticles. ACS Catal.5, 5673–5678 (2015). [Google Scholar]
- 5.Zecevic, J., Vanbutsele, G., De Jong, K. P. & Martens, J. A. Nanoscale intimacy in bifunctional catalysts for selective conversion of hydrocarbons. Nature528, 245–248 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.van Deelen, T. W., Hernández Mejía, C. & de Jong, K. P. Control of metal-support interactions in heterogeneous catalysts to enhance activity and selectivity. Nat. Catal.2, 955–970 (2019). [Google Scholar]
- 7.Gänzler, A. M. et al. Tuning the structure of platinum particles on ceria in situ for enhancing the catalytic performance of exhaust gas catalysts. Angew. Chem. Int. Ed.56, 13078–13082 (2017). [DOI] [PubMed] [Google Scholar]
- 8.Moliner, M. et al. Reversible transformation of Pt nanoparticles into single atoms inside high-silica chabazite zeolite. J. Am. Chem. Soc.138, 15743–15750 (2016). [DOI] [PubMed] [Google Scholar]
- 9.Jones, J. et al. Thermally stable single-atom platinum-on-ceria catalysts via atom trapping. Science353, 150–154 (2016). [DOI] [PubMed] [Google Scholar]
- 10.Xiong, H., Datye, A. K. & Wang, Y. Thermally stable single-atom heterogeneous catalysts. Adv. Mater.33, 2004319 (2021). [DOI] [PubMed] [Google Scholar]
- 11.Schwab, G.-M. Electronics of supported catalysts. Adv. Catal.27, 1–22 (1979). [Google Scholar]
- 12.Paredis, K. et al. Evolution of the structure and chemical state of Pd nanoparticles during the in situ catalytic reduction of NO with H2. J. Am. Chem. Soc.133, 13455–13464 (2011). [DOI] [PubMed] [Google Scholar]
- 13.Schweitzer, N. M. et al. High activity carbide supported catalysts for water gas shift. J. Am. Chem. Soc.133, 2378–2381 (2011). [DOI] [PubMed] [Google Scholar]
- 14.Yao, S. et al. Atomic-layered Au clusters on α-MoC as catalysts for the low-temperature water-gas shift reaction. Science357, 389 (2017). [DOI] [PubMed] [Google Scholar]
- 15.Li, Z. et al. Direct methane activation by atomically thin platinum nanolayers on two-dimensional metal carbides. Nat. Catal.4, 882–891 (2021). [Google Scholar]
- 16.Lin, L. et al. Low-temperature hydrogen production from water and methanol using Pt/α-MoC catalysts. Nature544, 80–83 (2017). [DOI] [PubMed] [Google Scholar]
- 17.Zhang, X. et al. Highly dispersed copper over β-Mo2C as an efficient and stable catalyst for the reverse water gas shift (RWGS) reaction. ACS Catal.7, 912–918 (2017). [Google Scholar]
- 18.Ge, Y. et al. Maximizing the synergistic effect of CoNi catalyst on α-MoC for robust hydrogen production. J. Am. Chem. Soc.143, 628–633 (2020). [DOI] [PubMed] [Google Scholar]
- 19.Li, S. et al. Atomically dispersed Ir/α-MoC catalyst with high metal loading and thermal stability for water-promoted hydrogenation reaction. Natl. Sci. Rev.9, 1–10 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Lin, L. et al. Atomically dispersed Ni/α-MoC catalyst for hydrogen production from methanol/water. J. Am. Chem. Soc.143, 309–317 (2021). [DOI] [PubMed] [Google Scholar]
- 21.Zhang, X. et al. A stable low-temperature H2-production catalyst by crowding Pt on α-MoC. Nature589, 396–401 (2021). [DOI] [PubMed] [Google Scholar]
- 22.Lu, Y. et al. Enhanced performance of xNi@yMo-HSS catalysts for DRM reaction via the formation of a novel SiMoOx species. Appl. Catal. B: Environ.291, 120075 (2021). [Google Scholar]
- 23.Liu, S., Wang, H., Smith, K. J. & Kim, C. S. Hydrodeoxygenation of 2-methoxyphenol over Ru, Pd, and Mo2C catalysts supported on carbon. Energy Fuels31, 6378–6388 (2017). [Google Scholar]
- 24.Wang, H., Liu, S., Govindarajan, R. & Smith, K. J. Preparation of Ni-Mo2C/carbon catalysts and their stability in the HDS of dibenzothiophene. Appl. Catal. A: Gen.539, 114–127 (2017). [Google Scholar]
- 25.Wang, H., Liu, S. & Smith, K. J. Synthesis and hydrodeoxygenation activity of carbon supported molybdenum carbide and oxycarbide catalysts. Energy Fuels30, 6039–6049 (2016). [Google Scholar]
- 26.Dong, J., Fu, Q., Jiang, Z., Mei, B. & Bao, X. Carbide-supported Au catalysts for water–gas shift reactions: a new territory for the strong metal-support interaction effect. J. Am. Chem. Soc.140, 13808–13816 (2018). [DOI] [PubMed] [Google Scholar]
- 27.Ma, Y. et al. High-density and thermally stable palladium single-atom catalysts for chemoselective hydrogenations. Angew. Chem. Int. Ed.132, 21797–21803 (2020). [DOI] [PubMed] [Google Scholar]
- 28.Cai, H. et al. Re-dispersion of platinum from CNTs substrate to α-MoC1-x to boost the hydrogen evolution reaction. Small19, 2207146 (2023). [DOI] [PubMed]
- 29.Diao, Y. et al. Plasma-assisted dry reforming of methane over Mo2C-Ni/Al2O3 catalysts: Effects of β-Mo2C promoter. Appl. Catal. B: Environ.301, 120779 (2022). [Google Scholar]
- 30.Zhang, X. et al. A novel Ni-MoCxOy interfacial catalyst for syngas production via the chemical looping dry reforming of methane. Chem9, 102–116 (2023). [Google Scholar]
- 31.Zhou, H. et al. Engineering the Cu/Mo2CTx (MXene) interface to drive CO2 hydrogenation to methanol. Nat. Catal.4, 860–871 (2021). [Google Scholar]
- 32.Lin, L. et al. Reversing sintering effect of Ni particles on γ-Mo2N via strong metal support interaction. Nat. Commun.12, 1–11 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Du, X. et al. Efficient catalytic conversion of jatropha oil to high grade biofuel on Ni-Mo2C/MCM-41 catalysts with tuned surface properties. J. Energy Chem.61, 425–435 (2021). [Google Scholar]
- 34.Peri, J. B. Computerized infrared studies of molybdenum/alumina and molybdenum/silica catalysts. J. Phys. Chem.86, 1615–1622 (1982). [Google Scholar]
- 35.Sundaramurthy, V., Dalai, A. K. & Adjaye, J. Effect of phosphorus addition on the hydrotreating activity of NiMo/Al2O3 carbide catalyst. Catal. Today125, 239–247 (2007). [Google Scholar]
- 36.Liang, G. et al. ZSM-5-supported multiply-twinned nickel particles: formation, surface properties, and high catalytic performance in hydrolytic hydrogenation of cellulose. J. Catal.325, 79–86 (2015). [Google Scholar]
- 37.Hadjiivanov, K. et al. Characterization of Ni/SiO2 catalysts prepared by successive deposition and reduction of Ni2+ ions. J. Catal.185, 314–323 (1999). [Google Scholar]
- 38.Moulijn, J. A., van Diepen, A. E. & Kapteijn, F. Catalyst deactivation: is it predictable?: what to do? Appl. Catal. A: Gen.212, 3–16 (2001). [Google Scholar]
- 39.Yao, S. et al. Exploring metal–support interactions to immobilize subnanometer Co clusters on γ-Mo2N: a highly selective and stable catalyst for CO2 activation. ACS Catal.9, 9087–9097 (2019). [Google Scholar]
- 40.Wei, H. et al. FeOx-supported platinum single-atom and pseudo-single-atom catalysts for chemoselective hydrogenation of functionalized nitroarenes. Nat. Commun.5, 1–8 (2014). [DOI] [PubMed] [Google Scholar]
- 41.Yu-Yao, Y.-F. & Kummer, J. T. Low-concentration supported precious metal catalysts prepared by thermal transport. J. Catal.106, 307–312 (1987). [Google Scholar]
- 42.Hao, Z. et al. Decoupling the effect of Ni particle size and surface oxygen deficiencies in CO2 methanation over ceria supported Ni. Appl. Catal. B: Environ.286, 119922 (2021). [Google Scholar]
- 43.Feng, K. et al. Experimentally unveiling the origin of tunable selectivity for CO2 hydrogenation over Ni-based catalysts. Appl. Catal. B: Environ.292, 120191 (2021). [Google Scholar]
- 44.Vogt, C. et al. Unravelling structure sensitivity in CO2 hydrogenation over nickel. Nat. Catal.1, 127–134 (2018). [Google Scholar]
- 45.Simons, J. F. M. et al. Structure sensitivity of CO2 hydrogenation on Ni revisited. J. Am. Chem. Soc.145, 20289–20301 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Lin, T. C., Bickel Rogers, E. E. & Bhan, A. Kinetic and thermodynamic considerations in thermocatalytic CO2 hydrogenation. ACS Catal.15, 780–788 (2025). [Google Scholar]
- 47.Vance, C. K. & Bartholomew, C. H. Hydrogenation of carbon dioxide on group viii metals: III, Effects of support on activity/selectivity and adsorption properties of nickel. Appl. Catal.7, 169–177 (1983). [Google Scholar]
- 48.Mutschler, R., Moioli, E., Luo, W., Gallandat, N. & Züttel, A. CO2 hydrogenation reaction over pristine Fe, Co, Ni, Cu and Al2O3 supported Ru: comparison and determination of the activation energies. J. Catal.366, 139–149 (2018). [Google Scholar]
- 49.Pu, T., Shen, L., Xu, J., Peng, C. & Zhu, M. Revealing the dependence of CO2 activation on hydrogen dissociation ability over supported nickel catalysts. AIChE J.68, 17458 (2022). [Google Scholar]
- 50.Li, J. et al. Direct conversion of cellulose using carbon monoxide and water on a Pt–Mo2C/C catalyst. Energy Environ. Sci.7, 393–398 (2014). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data generated in this study are available within the paper, its supplementary information files, and the repository. Source data are provided with this paper (Figshare 10.6084/m9.figshare.29924090). Data are available from the corresponding authors upon request.





