Open Access Article
Yao
Zhang
ab,
Huaizhong
Xiang
bc,
Boji
Wang
b,
Zhipeng
Qie
bd,
Keran
Jiao
b,
Xuzhao
Liu
f,
Xiaolei
Fan
*be and
Shanshan
Xu
*bf
aSchool of Civil Engineering, Weifang University of Science and Technology, Weifang 262700, China
bDepartment of Chemical Engineering, The University of Manchester, Oxford Road, Manchester M13 9PL, UK
cDepartment of Chemistry, Queen Mary University of London, London E1 4NS, UK
dCollege of Mechanical and Energy Engineering, Beijing University of Technology, Beijing 100124, China
eWenzhou Key Laboratory of Novel Optoelectronic and Nano Materials, Institute of Wenzhou, Zhejiang University, Wenzhou 325006, China
fDepartment of Materials, The University of Manchester, Oxford Road, Manchester M13 9PL, UK. E-mail: shanshan.xu@manchester.ac.uk
First published on 18th November 2025
Metal–support interaction (MSI) is a well-established strategy for tuning catalytic activity in thermal catalysis, yet its role in nonthermal plasma catalytic CO2 methanation remains insufficiently explored. In this study, Ni/CexZr1−xO2 catalysts were synthesized using CexZr1−xO2 supports calcined at different temperatures to systematically modulate the MSI. A volcano-shaped correlation was observed between the catalytic activity and support calcination temperature in both thermal and plasma systems. The CexZr1−xO2 support calcined at 600 °C having a moderate particle size, demonstrated the optimum MSI (i.e., promoting the facile formation of oxygen vacancies and stable interfacial anchoring of Ni particles) and thus the comparatively best catalytic performance under both conditions. Under the tested conditions, thermal CO2 methanation exhibited superior activity compared to plasma-assisted reactions, e.g., the NCZ-600 catalyst achieved an 83% CH4 yield at 350 °C versus 11.3% at 7.0 kV. These results underscore the critical role of the MSI in governing CO2 methanation across distinct catalytic environments and highlight its potential as a unifying design principle for both thermal and plasma catalysis.
MSI refers to the electronic and structural interactions occurring at the metal–oxide interface, which strongly influence catalyst activity, selectivity, and stability. These effects arise from thermodynamic driving forces that lead to electron redistribution between the metal and the oxide support to achieve electronic equilibrium, as well as from atomic migration that stabilizes surface structures such as overlayers or alloys.3 The phenomenon was first reported in 1978 by Tauster and co-workers,4 who observed suppressed CO and H2 chemisorption on group VIII metals supported on TiO2 following high-temperature reduction. They attributed this to a strong chemical interaction at the interface, later termed metal–support interaction (MSI). Subsequent studies revealed that similar interactions in reducible oxide-supported noble and transition metal catalysts could substantially alter catalytic performance, often via encapsulation of metal nanoparticles by partially reduced oxide species.5 In CO2 hydrogenation, it has been shown to regulate both conversion and product selectivity. An optimally tuned MSI can minimize surface energy, prevent nanoparticle aggregation, and thus enhance the adsorption and activation of intermediates (e.g., CO), thereby promoting further hydrogenation to methane.6–8 However, an excessively strong MSI may lead to over-encapsulation of active metal sites by the support, impeding reactant access and diminishing catalytic activity.6 Thus, carefully balancing the beneficial stabilization effects of the MSI against the risks of over-encapsulation is crucial for maximizing CO2 methanation efficiency and selectivity.
The conventional approach of modulating catalyst reduction temperature presents a fundamental trade-off. On the one hand, lowering the reduction temperature during catalyst preparation can partially alleviate over-encapsulation of active sites; on the other hand, it often results in insufficient electronic transfer and poor stabilization of metal nanoparticles.9–11 This strategy is typically effective only for precious metals, which are readily reduced at low temperatures, but it offers limited benefits for widely used transition metals such as nickel (Ni), the preferred choice for selective CO2 hydrogenation to methane.9 To address this limitation, an alternative strategy is to adjust the calcination temperature of the catalyst support prior to metal loading. This simple yet powerful approach allows systematic tuning of key support properties, such as morphology, crystallinity, and surface chemistry,12,13 thereby optimizing the MSI during subsequent metal deposition. Crucially, this method can enhance catalyst stability and activity while minimizing the risk of over-encapsulation.
In addition to catalyst design, nonthermal plasma (NTP) activation is another effective strategy for enabling CO2 methanation under mild conditions. NTP catalytic systems operate under ambient temperature and atmospheric pressure, while the energetic electrons in plasma typically possess energies of 1–10 eV, sufficient to activate the highly stable CO2 molecule.14–16 Consequently, compared with conventional thermal catalytic systems, NTP catalytic CO2 methanation can, in principle, achieve effective activation under significantly milder conditions. However, catalyst design strategies tailored specifically for NTP catalysis remain underexplored. For example, while tuning the metal–support interaction (MSI) is a well-established strategy in thermal catalysis, its applicability to NTP-assisted CO2 methanation has received limited attention.17 Considering the complex nature of plasma systems, where energetic electrons interact with ionized gases to generate a variety of reactive species such as radicals and ions, facilitating bond cleavage and formation, the role of the MSI in NTP systems may differ significantly from that in thermal catalysis.16 Therefore, the feasibility of applying MSI-based catalyst design in NTP-assisted CO2 methanation requires careful evaluation. Several pioneering studies have demonstrated the potential of this approach. For example, Sun et al. found that a plasma catalytic system using Pd/ZnO achieved nearly 36.7% CO2 conversion and 35.5% CO yield, significantly outperforming plasma-only or plasma + ZnO systems. This enhancement was attributed to ZnOx overlayers generated through strong MSI at the Pd–ZnO interface, combined with abundant hydrogen species on the Pd/ZnO surface.18 Similarly, our recent work demonstrated that modulating the MSI in Ru/ZrO2 catalysts, by tuning the ZrO2 phase composition, influenced CO2 hydrogenation under plasma conditions. Ru supported on mixed-phase ZrO2 achieved 81.3% CO2 conversion with 97.3% CH4 selectivity, whereas Ru on monoclinic or tetragonal ZrO2 promoted CO formation (>90%) with limited CO2 conversion (<14%).19 Despite these advances, comparative studies that systematically probe the influence of MSI on CO2 methanation under both thermal and NTP conditions are scarce, and the underlying mechanisms of such design strategies remain poorly understood.
In this study, we proposed a simple method (i.e., varying calcination temperature of the CexZr1−xO2 support, CZ) to tune the particle size of the CZ support and thereby regulate the MSI in the resulting Ni/CexZr1−xO2 (NCZ) catalysts. We systematically evaluated the catalytic performance of these NCZ catalysts for CO2 methanation under both thermal and plasma conditions and correlated activity with structural and electronic properties. Among the series, the CZ support calcined at 600 °C (CZ-600) afforded the most favorable MSI, enabling stable anchoring of Ni particles and promoting oxygen vacancy formation, which facilitated CO2 adsorption and activation. As a result, NCZ-600 delivered the best catalytic performance in both thermal and NTP systems. Notably, while thermal catalysis achieved higher CH4 yields than NTP under the conditions studied, our results highlight the central role of the MSI in governing catalyst performance across both environments and point to additional factors unique to plasma catalysis that warrant further investigation.
:
HNO3
:
HF is 1
:
3
:
1) via the method of microwave digestion (ETHOS UP microwave digester). The hydrogen temperature-programmed reduction (H2-TPR) and CO2 temperature-programmed desorption (CO2-TPD) were performed on a Belcat II (Microtrac MRB, Japan) instrument. For H2-TPR, ∼50 mg of the calcined sample was reduced in 5% H2/Ar gas flow at 10 °C min−1 from 200 to 800 °C. For CO2-TPD, ∼50 mg of the calcined catalysts were reduced in situ at 500 °C in 10% H2/Ar and cooled down to 60 °C. CO2 gas flow with the flowrate of 30 mL min−1 was purged for 1 h with the aim of adsorption. The physically adsorbed CO2 was then removed by purging He for 1 h (at a flow rate of 30 mL min−1). A mass spectrometer (MS, Hiden™ HPR-20) was used to analyze the effluent gas as the temperature increased from 60 to 800 °C with a ramping rate of 30 mL min−1. The morphology of reduced NCZ catalysts and size of CZ supports were obtained by high-resolution transmission electron microscopy (HRTEM) at 300 kV on a Tecnai F30 (FEI, Germany) electron microscope equipped with a cold field emission gun (FEG) source. The scanning transmission electron microscope (STEM) and energy-dispersive X-ray spectroscopy (EDX) mapping images were obtained with a Titan STEM (G2 80-200), equipped with a high angle angular dark field (HAADF) detector. Each powder sample was dissolved and dispersed in ethanol and then drop-casted on a 300-mesh holey carbon grid prior to imaging.
:
1) balanced with Ar at a constant weight hourly space velocity (WHSV) of 30
000 mL (STP) gcat−1 h−1 was used for the reaction. The activity tests were carried out at between 100 and 500 °C with an interval of 50 °C. The composition of the outlet gas was monitored on-line by MS (Hiden™ HPR-20).
NTP-catalytic CO2 methanation was carried out in a dielectric barrier discharge (DBD) plasma reactor (Fig. S2). Details of the NTP reactor system are described elsewhere.2 The catalysts were all reduced under the same conditions above. However, due to the sample exposition to air, the reduced catalysts (∼200 mg of pelletized catalysts, 355–500 μm) were treated again in situ by reducing plasma discharge (in H2 balanced by Ar flow) at 6.5 kV for 0.5 h. Subsequently, a CO2/H2 gas flow with a molar ratio of 1
:
4 diluted in Ar was introduced into the NTP reactor for the reaction at a WHSV of 30
000 mL (STP) gcat−1 h−1. The applied voltage adopted a range of 5.5–7.0 kV with a constant frequency of 20.5 kHz.
The CO2 conversion (XCO2), selectivity to CH4 (SCH4) and CO (SCO), yield of CH4 (YCH4) and carbon balance (Cbalance) are defined based on eqn (1)–(5):
![]() | (1) |
![]() | (2) |
![]() | (3) |
| YCH4 = XCO2 × SCH4 | (4) |
![]() | (5) |
![]() | (6) |
For plasma catalysis, the apparent energy barrier (Ea,NTP, kJ mol−1) was estimated following the approach reported in the literature, as shown in eqn (7):21
![]() | (7) |
Using these methods, we determined the apparent activation energies of the catalysts under both thermal and plasma conditions (Fig. S5 and Table S1).
| Catalyst | Nia (wt%) | N2 physisorption | d Ni c (nm) | Dimension | XPS analysis | |||||
|---|---|---|---|---|---|---|---|---|---|---|
| S BET b (m2 g−1) | V p b (cm3 g−1) | D p b (nm) | d Ni d (nm) | d NiO e (nm) | D support f (nm) | Ce3+/(Ce3+ + Ce4+)g (%) | Osurface/(Olattice + Osurface + Oothers)h (%) | |||
| a Determined by ICP-OES. b By N2 physisorption. c Average particle size of Ni on the reduced NCZ-T catalysts calculated by the Debye–Scherrer equation based on XRD patterns. d Average particle size of Ni on the reduced catalysts from HRTEM images (Fig. 2). e Average particle size of NiO on the calcined NCZ-T catalysts based on XRD results. f Average support size of the reduced NCZ-T catalysts, obtained from TEM images (Fig. S4). g Extracted from the deconvoluted Ce 3d XPS spectra. h Estimated oxygen vacancy concentration from O 1s XPS spectra. | ||||||||||
| NCZ-500 | 15.2 | 30 | 0.066 | 8.4 | 4.1 | 4.1 | 3.6 | 15.1 | 27 | 46.5 |
| NCZ-600 | 15 | 22 | 0.065 | 10.3 | 4.8 | 6.7 | 3.6 | 13.8 | 24.1 | 50.7 |
| NCZ-700 | 14.9 | 21 | 0.065 | 12.5 | 6.6 | 7.2 | 3.8 | 17.2 | 26.2 | 43.9 |
| NCZ-800 | 15 | 18 | 0.064 | 14.1 | 7.3 | 8.2 | 3.9 | 18.9 | 20 | 40.7 |
| NCZ-900 | 15 | 10 | 0.051 | 18.8 | 7.9 | 9.8 | 3.8 | 24.8 | 17.4 | 36.2 |
The volcano-shaped activity trend observed with respect to the support calcination temperature is likely attributed to changes in the support size rather than the Ni particle size (Fig. 1 and S8 and Table 2). Specifically, similar average NiO particle sizes (∼3.6–3.9 nm) across all calcined NCZ-T samples were observed. However, after reduction at 500 °C, the metallic Ni particle size increases significantly from ∼4.1 nm in NCZ-500 to ∼9.8 nm in NCZ-900 (Table 2, Fig. 2, S6 and S8). Interestingly, the NCZ-500 catalyst, which has the smallest Ni particle size (i.e., the highest metal surface area), does not show favorable CO2 methanation activity, indicating that the Ni particle size is not the primary factor to affect the catalyst performance.26 Instead, NCZ-600 with an optimal support structure, which demonstrates the highest activity, does not exhibit the smallest Ni particles or the largest Ni surface area. This suggests that the enhanced catalytic performance of the NCZ-600 catalyst is primarily attributed to the strength of the metal–support interaction between Ni nanoparticles and CZ-T supports, rather than Ni particle size effects.
![]() | ||
| Fig. 2 TEM images and corresponding support size of the reduced (a and a′) NCZ-500 (b and b′) NCZ-600 (c and c′) NCZ-700 (d and d′) NCZ-800 and (e and e′) NCZ-900 catalysts (reduced at 500 °C). | ||
After reduction at 500 °C, metallic Ni particle sizes increased progressively from ∼4.1 to 9.8 nm across the catalyst series, as determined by both STEM and XRD (Fig. S8), suggesting Ni sintering/agglomeration at higher calcination temperatures. This agglomeration is likely a consequence of the decreased surface area of the CZ supports at elevated calcination temperatures (e.g., SBET = 10 m2 g−1 for NCZ-900, Table 2). Interestingly, TEM results (Tables 1 and 2) show that the NCZ-600 catalyst retained a support size (∼13.8 nm) comparable to that of the parent CZ-600 support after reduction. Conversely, other catalysts exhibited a pronounced increase in the support size after reduction (Fig. S9 and 2). This finding suggests that NCZ-600, possessing an intermediate support size, offers an optimal balance of MSI and thermal stability, effectively suppressing support surface area loss during high-temperature treatment. Such stabilization minimizes Ni sintering and preserves active interfacial sites, which accounts for the superior catalytic activity of NCZ-600 under both thermal and plasma conditions.
In addition to providing stable anchoring sites for Ni species, modulation of the MSI also influences the reductivity of Ni species on the support. To investigate this, temperature-programmed reduction with hydrogen (H2-TPR) experiments were conducted. As shown in Fig. 3, four reduction peaks were observed for all catalysts. The lowest-temperature peak (α), appearing below ∼300 °C, corresponds to the removal of surface adsorbates or the reduction of easily reducible surface Ni2+ species. The highest-temperature small peak, above ∼700 °C, is attributed to the bulk reduction of Ce4+ species, which is relatively unaffected by the presence of Zr or Ni species.27 The dominant peak β, centered around ∼366 °C, can be ascribed to the reduction of bulk or subsurface Ni2+ species. Additionally, a smaller shoulder peak γ (∼410–453 °C) is associated with the reduction of Ce4+ to Ce3+ within the distorted fluorite structure, accompanied by the formation of oxygen vacancies, and is influenced by the presence of dispersed Ce4+ and Ni2+ ions.27,28
For NCZ-600, the Ni2+ reduction peak (β) appears at a slightly lower temperature (∼356 °C) compared with other catalysts, before shifting progressively to higher temperatures as the support calcination temperature increases to 900 °C. This behavior suggests that NCZ-600 exhibits the weakest MSI within the series, facilitating Ni2+ reduction. By contrast, the decreased CO2 methanation activity of NCZ-800 and NCZ-900 is likely due to the stronger MSI, which encouraged the formation of a thicker support overlayer over the Ni particles.29 Meanwhile, the γ peak displays a systematic shift toward lower temperatures (from 453 °C to 410 °C) with increasing support calcination temperature (500 °C to 900 °C), reflecting enhanced reducibility of Ce4+ in the distorted fluorite structure. Notably, H2 consumption for the γ peak reaches its maximum in NCZ-600, indicating the highest concentration of oxygen vacancies achieved after calcination at 600 °C. After calcination at higher temperatures (≥700 °C), the γ peak decreases significantly, suggesting a reduced extent of Ce4+ to Ce3+ transformation and consequently fewer oxygen vacancies.
![]() | ||
| Fig. 4 XRD patterns of (a) the reduced NCZ-T catalysts, (b) the calcined NCZ-T samples and (c) the calcined CZ-T supports. | ||
Raman spectroscopy was subsequently employed to investigate the oxygen vacancy formation (Fig. 5a). The phase composition of the zirconia–ceria system can form distinct structural phases depending on the calcination temperature and the Zr/Ce ratio.31 According to the literature, the strong broad band at 467 cm−1, along with two weak bands at ∼260 cm−1 (Eg), and 315 cm−1 (B1g), originates from the six active Raman modes of A1g + 2B1g + 3Eg associated with the tetragonal ZrO2 phase (space group P42/nmc), indicating the presence of a small amount of tetragonal ZrO2 in the samples.32,33 The prominent Raman peak at ∼467 cm−1, observed in all samples, is attributed to the F2g vibrational mode of the fluorite cubic structure of CeO2 (symmetrical stretching of the Ce–O bonds), which is associated with the presence of oxygen vacancies. These vacancies are formed through the substitution of Ce4+ by Zr4+, leading to the generation of a CeO2–ZrO2 solid solution in the CZ-T supports, which is consistent with the above XRD results. In the reduced NCZ-T catalysts, additional oxygen vacancies are introduced due to the partial reduction of Ce4+ to Ce3+ during the reduction process.27,34 Furthermore, a broad and weak band at ca. 620 cm−1 in both CZ-T and NCZ-T samples, namely the defect-induced longitudinal optical band (i.e., LO band), can be assigned to structural defects and oxygen vacancies within the fluorite lattice.27,35
![]() | ||
| Fig. 5 Raman spectra for (a) the CZ-T supports and (b) the reduced NCZ-T catalysts and (c) the corresponding ILO/IF2g values. | ||
The intensity ratio of the LO to the F2g bands can be used as an indicator of the oxygen vacancy density.27 As shown in Fig. 5(c), this ratio shows no significant variation among the CZ-T supports or the calcined NCZ-T samples, suggesting that the oxygen vacancy concentration remains relatively unchanged after calcination. Interestingly, after reduction treatment at 500 °C, the ILO/IF2g ratio for the reduced NCZ-T catalysts initially increased and then decreased, displaying a volcano-like trend with NCZ-600 showing the highest value of 0.20 (Fig. 5c). This suggests that the variation in oxygen vacancy density is not from the calcination of the supports or the loading of nickel species, but rather from the reduction process, specifically the partial reduction of Ce4+ to Ce3+ in the CenO2n-based lattice during catalyst activation.36 The highest oxygen vacancy density observed in NCZ-600 can thus be attributed to an optimal MSI, where Ni nanoparticles interact with supports to promote partial reduction of Ce4+ to Ce3+. This enhanced generation of Ce3+ leads to a higher concentration of oxygen vacancies, which in turn facilitates superior catalytic performance.
To further elucidate the MSI and its influence on the formation of oxygen vacancies, the surface characteristics of all NCZ-T catalysts were examined using X-ray photoelectron spectroscopy (XPS) analysis (Fig. S12a and 6). In the O 1s spectrum, the photoelectron peaks correspond to: (i) lattice oxygen (Olattice) at ∼529.4 eV, (ii) surface oxygen species (Osurface: O2−, O22− or O−) at ∼531.1 eV and (iii) weakly bound oxygen species such as CO32−, hydroxyl groups and surface-adsorbed H2O (Oothers) at ∼533.0 eV.37 As shown in Fig. 6 and Table 2, the concentration of oxygen vacancies, calculated as Osurface/Olattice + Osurface + Oothers, increased initially and then decreased again with increasing support calcination temperature, with maximum oxygen vacancy concentration achieved by NCZ-600, which is consistent with the results from H2-TPR (γ peak) and Raman measurements (ILO/IF2g).
This phenomenon is further supported by the Ce 3d XPS spectra, which display two sets of multiplets corresponding to the spin–orbit components 3d3/2 and 3d5/2. The deconvoluted peaks labeled v0 (879.91 eV), v2 (884.48 eV), u0 (898.55 eV) and u2 (902.84 eV) are assigned to the Ce 3d104f1 state of Ce3+, while the peaks labeled v1 (882 eV), v3 (888.4 eV), v4 (897.8 eV), u1 (900.7 eV), u3 (907.2 eV) and u4 (917 eV) correspond to the Ce 3d104f0 state of Ce4+.27 A decreasing Ce signal is observed with increasing calcination temperature of the supports, indicating aggregation of Ce species in catalysts with supports calcined at higher temperatures (Fig. S12b and 6b). CeO2 has the capacity to store and release oxygen through the redox reaction (2CeO2 ↔ Ce2O3 + 1/2O2), where each released oxygen atom creates an oxygen vacancy and simultaneously converts two Ce4+ ions to Ce3+via electron transfer.36 Thereby, the ratio of Ce3+/(Ce3+ + Ce4+) can be used to indicate the oxygen vacancy density. This ratio follows the same trend as the concentration of oxygen vacancies observed in the O 1s spectra, further confirming the consistency of the findings.
Based on above findings, we can conclude that the support calcination temperature of 600 °C, associated with weak metal support interaction, promotes the formation of oxygen vacancies. The facile generation of these vacancies is beneficial for CO2 methanation, where the role of the support is critical.38 The CZ-T supports are particularly effective in CO2 adsorption and its activation at the interface, owing to their pronounced redox properties and abundant oxygen vacancies, which can readily interact with CO2 molecules to generate reactive intermediates such as carboxylates during CO2 methanation.13,39 This enhancement in CO2 adsorption due to oxygen defects is further supported by CO2-TPD results (Fig. S13), where NCZ-600 shows moderate basic sites for CO2 adsorption. In plasma-assisted systems, where plasma discharges can create new active sites, or reactivate adsorption sites, tuning the MSI may potentially enable new reaction pathways by modulating oxygen vacancy formation under plasma conditions.40–42 This may also explain the comparatively lower catalytic activities observed in plasma environments than that of thermal catalysis when applying the same MSI-based strategy. Previous studies have shown that plasma-catalytic CO2 methanation over supported Ni catalysts proceeds via similar pathways to thermal catalysis, that is, forming surface carbonate and/or formate intermediates, which are subsequently hydrogenated to methane.41–43 The key difference in plasma-catalytic CO2 methanation is the activation of reactant molecules in gas discharge by electron-impact dissociation and/or gas-phase reactions. For example, plasma can activate CO2 and H2 in the gas discharge to form reactive intermediates such as vibrationally excited CO2, CO, and H species. These short-lived gas species can directly interact with catalyst surfaces and participate in the hydrogenation via Hinshelwood–Rideal mechanisms, forming a CO-hydro pathway for methane, as demonstrated by previous studies.43,44
The relevant post-plasma characterisation, including Raman and XRD analyses, was conducted to evaluate possible surface modifications and MSI changes (Fig. S10 and S11, Pages S9). As shown by XRD in Fig. S10, negligible changes were observed in the patterns of the used NCZ-T samples, and no NiO diffraction peaks were detected, confirming the absence of Ni oxidation and indicating similar Ni dispersions compared to the fresh catalyst under plasma discharge conditions. Raman results further revealed that the abundance of oxygen vacancies (reflected by ILO/IF2g values) in the used NCZ-T catalysts had the same volcano-type trend as in the reduced catalysts (with NCZ-600 showing the maximum concentration), suggesting that the MSI features of the catalysts under investigation were preserved after plasma catalysis. A slight overall increase in oxygen vacancy intensity was observed for all used catalysts, being similar to previous findings that plasma treatment can generate additional oxygen vacancies via oxygen atom removal from the CeO2 surface.45–47 Overall, these results suggest that plasma discharge does not cause significant changes on the Ni state and MSI during the plasma-catalytic reactions.
Supplementary information: the supplementary activity results, BET, STEM, XPS and characterisation of spent catalysts are included in the SI. See DOI: https://doi.org/10.1039/d5cy00847f.
| This journal is © The Royal Society of Chemistry 2026 |