Next Article in Journal
Halloysite@Polydopamine Nanoplatform for Ultrasmall Pd and Cu Nanoparticles: Suitable Catalysts for Hydrogenation and Reduction Reactions
Previous Article in Journal
Interfacial Stabilization Strategy: Hydrothermally Synthesized Highly-Dispersed and Low-Leaching CuO-Biochar for Efficient Peroxydisulfate Activation and Cu-EDTA Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unraveling the Reaction Mechanism of the Reverse Water–Gas Shift Reaction over Ni/CeO2 and CeO2−x Catalysts

1
State Key Laboratory of Heavy Oil Processing, College of Chemistry and Chemical Engineering, China University of Petroleum (East China), Qingdao 266580, China
2
National Institute of Clean-and-Low-Carbon Energy, Future Science City, Changping District, Beijing 102211, China
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1028; https://doi.org/10.3390/catal15111028 (registering DOI)
Submission received: 10 October 2025 / Revised: 26 October 2025 / Accepted: 28 October 2025 / Published: 1 November 2025

Abstract

The reverse water–gas shift (RWGS) reaction efficiently converts CO2 to CO, with vital applications in carbon emission reduction and Fischer-Tropsch chemical production. This study used density functional theory (DFT) to investigate CO2 adsorption and activation on CeO2, oxygen-vacancy CeO2 (CeO2−x), and single-atom Ni-loaded CeO2 (Ni/CeO2). Adsorption energy analysis indicates that CO2 preferentially adsorbs at the intermediate oxygen sites on CeO2 and Ni/CeO2, but on CeO2−x, it preferentially adsorbs at the oxygen vacancies. Mulliken charge and band gap results indicate that CeO2−x and Ni/CeO2 exhibit higher activity than pure CeO2. Density of states studies indicate that CeO2, CeO2−x, and Ni/CeO2 can activate CO2 to varying degrees; strong hybridization between Ni’s d-orbitals and CO2’s O p-orbitals is key to Ni/CeO2’s high activity. Mechanistically, CeO2−x follows the RWGS redox mechanism, while Ni/CeO2 follows the formate-associated mechanism. This work innovatively clarifies differential CO2 adsorption-activation by vacancies and Ni in CeO2-based catalysts, providing a theoretical basis for RWGS catalyst design and supporting low-energy carbon conversion development.

1. Introduction

CO2, as a major component of greenhouse gases, primarily originates from the burning of fossil fuels and has caused serious climate and environmental problems [1]. In this context, controlling excessive CO2 emissions through carbon capture, storage, and utilization (CCUS) has attracted great attention from countries around the world [2,3]. Converting the well-known greenhouse gas CO2 into chemical fuels is a very promising approach, as it not only helps mitigate global warming but also serves as an alternative to chemical fuels [4]. In CO2 conversion technology, catalytic hydrogenation of CO2 is considered one of the most commercially feasible methods [5], especially by converting CO2 into CO through the endothermic RWGS reaction. Since CO is an important chemical raw material for the synthesis of fine chemicals and fuels, converting CO2 into CO through the RWGS reaction can obviously bring considerable economic benefits and provide a reliable solution to energy crises and environmental challenges [6,7].
However, CO2 is a molecule with high kinetic and thermodynamic stability, as the cleavage of its C=O bonds requires a high activation barrier under ambient conditions. Under normal conditions, the cleavage of the C=O bond in CO2 molecules requires a very high activation barrier, making the development of efficient catalysts urgently needed. A good catalyst should have a low hydrogenation activation barrier, good C=O bond cleavage ability, and the ability to dissociate hydrogen [8]. In order to obtain high-performance RWGS reaction catalysts, researchers have conducted extensive studies. Research on noble metal catalysts (such as Pt [9], Au [10], and Pd [11]) has found that they all possess suitable hydrogenation capabilities and good RWGS catalytic activity. However, noble metals are prone to sintering at high temperatures, resulting in poor stability of noble metal catalysts, high costs, and low natural abundance, which is unfavorable for their widespread application in industrial production. To reduce costs and improve the high-temperature stability of the corresponding catalysts, researchers have started studying transition metal catalysts [12]. Transition metal catalysts, especially nickel, are widely used in CO2 hydrogenation reactions due to their advantages such as low cost and high activity at low temperatures, including CO2 methanation, CO2 hydrogenation to methanol or formic acid, and water–gas shift reactions [13,14]. Although nickel catalysts exhibit high hydrogenation activity, they face issues such as carbon deposition and sintering. To address these problems, the oxide supports of many RWGS catalysts have been studied, including CeO2, Al2O3, SiO2, TiO2, and ZrO2 [15]. CeO2 is known for its unique redox properties because it can rapidly interconvert between Ce4+ and Ce3+ [16]. At the same time, Ce3+ ions can enhance the stability of the CeO2 lattice, helping to prevent carbon deposition on the catalyst surface [17]. In addition, the high oxygen mobility on the surface of CeO2 also promotes the activation and transfer of oxygen molecules, which is crucial for the catalytic process. The presence of a large number of oxygen vacancies on nickel catalysts supported on reducible oxides originates from the metal-support interaction, which facilitates the adsorption and activation of CO2 [18,19]. Nickel catalysts supported on CeO2 exhibit excellent catalytic performance in terms of activity, selectivity, and stability in the RWGS reaction [20]. However, most nickel catalysts exhibit high CH4 selectivity in the hydrogenation of CO2. Therefore, developing catalysts for the RWGS reaction that are highly active, highly selective, and low-cost remains a challenge.
Numerous studies have shown that particle sizes significantly affect the selectivity of the RWGS reaction [21]. For the CO2 hydrogenation reaction, larger nickel particles are conducive to promoting the methanation of CO2, whereas smaller nickel particles are more favorable for the production of CO [22,23]. Improving the dispersion of nickel by reducing its loading, introducing a second metal additive to form bimetallic alloys, and activating metal-support interactions will be an effective strategy to enhance CO selectivity [24,25,26]. Wu et al. [27] studied nickel metals supported on silica with mass fractions ranging from 0.5% to 10%. Their findings indicate that different nickel loadings result in varying coverages of H species on the catalyst surface, thereby affecting the selectivity of CO and CH4. A similar trend has also been observed in Ni/Al2O3, Ni/MgO, and Ni/CeO2 catalysts, where a low Ni loading is associated with an increase in RWGS activity [28]. Christopher and others found that atomically dispersed isolated metal centers are favorable for the RWGS reaction, whereas metal clusters serve as active centers for carbon dioxide methanation [29]. DFT has become an important tool for studying the surface structure, active sites, and microscopic mechanisms of catalysts, providing a powerful method for elucidating the microscopic characteristics of catalysts [30]. Hu et al. [31] used DFT to simulate single Pt atoms loaded on TiO2(110) and TiO2(101), respectively, and found that single Pt loaded on TiO2(101) exhibited a lower energy barrier for COOH intermediate formation and a more favorable redox process, thereby enhancing the CO2 conversion rate. Yang et al. [32] found through experiments and DFT simulations that the RWGS reaction catalyzed by single-atom Re1 loaded on TiO2 follows the formate pathway, while the RWGS reaction catalyzed by Re7 clusters loaded on TiO2 follows the redox pathway. Li et al. [33] demonstrated through experiments and DFT calculations that atomically dispersed Co-N4 centers follow a hydrogen-assisted pathway, where the intermediate COOH* desorbs and dissociates into CO, while Co nanoparticle catalysts mainly follow a direct dissociation pathway. Although some progress has been made in highly efficient RWGS reaction catalysts, the rate-determining steps and mechanisms followed in the RWGS reaction catalyzed by CeO2−x and Ni/CeO2 have not been fully clarified, which limits further optimization and application of the catalysts.
The purpose of this study is to develop catalysts with high activity, high selectivity, and high stability. In this study, we investigated the effects of CeO2, CeO2−x, and Ni/CeO2 on CO2 catalysis using Mulliken charges and band gaps in DFT. By analyzing the adsorption energies and geometric parameters of CO2 on different CeO2 surfaces, we identified the optimal adsorption sites for CO2 on each surface. Next, to explore the influence of different CeO2 on CO2 orbitals, we analyzed the density of states of CO2 on CeO2, CeO2−x, and Ni/CeO2. Finally, to investigate the rate-determining steps and reaction mechanisms of CeO2−x and Ni/CeO2 catalyzed RWGS reactions, we conducted transition state searches. Through DFT calculations, this study provides theoretical guidance for designing efficient CeO2-based RWGS reaction catalysts.

2. Results and Discussion

2.1. Mulliken Charges on CeO2, CeO2−x, and Ni/CeO2

The Ce and O atoms in different CeO2 and their Mulliken charges are shown in Figure 1 and Table 1. As can be seen from Table 1, compared with pure CeO2, the Mulliken charges of Ce1, O1, and O3 atoms on the CeO2−x surface increased by 0.147 |e|, −0.06 |e|, and −0.06 |e|, respectively, after introducing oxygen vacancies. This indicates that the introduction of oxygen vacancies significantly alters the charge distribution of Ce1, O1, and O3 atoms, which is conducive to enhancing the catalytic activity of CeO2(111). After introducing a single Ni atom, the changes in the Mulliken charges of Ce1, Ce2, and Ce3 atoms in CeO2 are 0.001|e|, 0.187|e|, and −0.025|e|, respectively. The Mulliken charge of the Ce2 atom connected to Ni changes the most, indicating a significant transfer of charge. This indicates that the introduction of single-atom Ni significantly changes the charge distribution of Ce2 atoms, thereby helping to enhance the catalytic activity of CeO2.

2.2. Optimal Adsorption Sites of CO2 on CeO2, CeO2−x, and Ni/CeO2

To explore the optimal adsorption sites of CO2 on different CeO2 surfaces, we investigated the adsorption of CO2 on the intermediate oxygen sites, edge oxygen sites, oxygen vacancies of CeO2 and CeO2−x, and Ni atoms of Ni/CeO2. The stable configurations, adsorption energies, and geometric parameters of different CO2 molecules adsorbed on different CeO2 surfaces are shown in Figure 2 and Table 2. As can be seen from Figure 2 and Table 2, for CO2 adsorption on the oxygen sites in the intermediate and at the edges of CeO2, the adsorption energies are −0.789 eV and −0.776 eV, respectively, indicating that CO2 adsorption on the intermediate oxygen sites of CeO2 has a higher adsorption energy and is more stable. CO2 adsorbs on the intermediate oxygen and edge oxygen sites of CeO2, with the longest C=O bond lengths being 1.232 and 1.227 Å, respectively. The C=O bond length is longest when CO2 is adsorbed on the intermediate oxygen site of CeO2, indicating that CO2 is most easily cleaved into CO at this site, which is favorable for CO2 conversion. CO2 is adsorbed on the intermediate oxygen of CeO2, and the distance between the carbon in CO2 and the oxygen in CeO2 is the shortest, at 1.405 Å, indicating that CO2 adsorption at the intermediate oxygen site of CeO2 is the most stable. For CO2 adsorption on CeO2−x, the adsorption energy of CO2 at oxygen vacancies of CeO2−x is higher than at edge oxygen sites, and the C=O bond length at the oxygen vacancies is the longest, indicating that CO2 is more easily adsorbed and dissociated into CO at the oxygen vacancies of CeO2−x. This is consistent with the oxygen vacancies formed in CeO2 serving as active centers for the activation of CO2 molecules [34]. For Ni/CeO2, we investigated three possible sites for CO2 adsorption: the oxygen positions in the middle of CeO2, the edge oxygen positions, and the Ni sites. From the adsorption energy, the C=O bond length, and the distance from the carbon in CO2 to various sites, it can be seen that CO2 has the highest adsorption energy, the longest C=O bond length, and the shortest distance from the carbon to the site at the intermediate oxygen position on Ni/CeO2. This indicates that the optimal adsorption site for CO2 on Ni/CeO2 is not on the Ni site, but on the oxygen site in the intermediate of Ni/CeO2. This is consistent with most of the CO2 dissociating on CeO2 rather than on Ni [35].

2.3. Band Gaps of CeO2, CeO2−x, and Ni/CeO2 Before and After CO2 Adsorption

In order to more accurately analyze the electronic structure characteristics of CeO2, CeO2−x, and Ni/CeO2, the band structures of different CeO2 before and after CO2 adsorption were compared, as shown in Figure 3. The reduction in the band gap decreases the energy required for electron transition from the valence band to the conduction band, which facilitates more electrons participating in surface catalytic reactions. After the introduction of oxygen vacancies or Ni loading, the band gaps of CeO2−x and Ni/CeO2 are reduced by 1.877 eV and 1.759 eV, respectively, compared to CeO2. Therefore, CeO2−x and Ni/CeO2 exhibit higher reactivity than CeO2. Before and after CO2 adsorption, the band gap of CeO2 changed by 0.05 eV, indicating that the electron distribution of CeO2 is hardly affected by CO2 adsorption. CeO2−x and Ni/CeO2 exhibit greater bandgap changes before and after CO2 adsorption, increasing from 0.094 eV to 0.837 eV and from 0.212 eV to 0.837 eV, respectively, indicating that CeO2−x and Ni/CeO2 interact more strongly with CO2 and can stably adsorb CO2 on their surfaces.

2.4. Density of States of CO2 Adsorbed on CeO2, CeO2−x, and Ni/CeO2

Figure 4 shows the density of states of CO2 adsorbed on CeO2, CeO2−x, and Ni/CeO2, where the C atom refers to the carbon atom in CO2, and Oa and Ob represent the two oxygen atoms in CO2, respectively. From Figure 4a,d,g, it can be seen that when CO2 is adsorbed on CeO2, CeO2−x, and Ni/CeO2, the orbitals of carbon in CO2 are different, indicating that the degree of CO2 activation is different on CeO2, CeO2−x, and Ni/CeO2. From Figure 4b,c, it can be seen that CO2 is adsorbed on CeO2, and the orbitals of Oa and Ob in CO2 are the same, indicating that CeO2 affects both oxygen atoms in CO2 equally. Similarly, from Figure 4e,f, it can be seen that CO2 is adsorbed on CeO2−x, and the orbitals of Oa and Ob in CO2 are the same, indicating that CeO2−x affects both oxygen atoms in CO2 equally. From Figure 4h,i, it can be seen that when CO2 is adsorbed on Ni/CeO2, the orbitals of Oa and Ob in CO2 are different. This may be due to the different distances between the oxygen atoms in CO2 and the Ni atoms, indicating that CO2 is activated in a highly asymmetric manner. At −5 eV and 0.5 eV, the d orbitals of the Ni atoms overlap with the p orbitals of the O atoms in CO2, indicating strong hybridization between the d orbitals of Ni and the p orbitals of the O atoms in CO2. This fully demonstrates that the interaction between CO2 and Ni atoms is mainly attributed to the hybridization between the d orbitals of Ni and the p orbitals of O in CO2.

2.5. Mechanism of the RWGS Reaction Catalyzed by CeO2−x and Ni/CeO2

There are several mechanisms for the RWGS reaction on CeO2-based catalysts, including the direct dissociation mechanism (redox mechanism) and the hydrogen-assisted mechanism (formate-associated mechanism and carboxylate-associated mechanism), the latter involving intermediates such as formates and carboxylates [36]. The basic steps of each mechanism of the RWGS reaction are shown in Table 3.

2.5.1. Redox Mechanism of CeO2−x Catalyzed RWGS Reaction

The redox mechanism of the CeO2−x catalyzed RWGS reaction is shown in Figure 5. CO2 adsorbs on CeO2−x, undergoing significant deformation, indicating that CO2 is activated, forming CO2*. Then H2 molecules adsorb onto CeO2−x and are dissociated into two H* atoms. CO2* is further activated by CeO2−x, breaking one C=O bond in CO2* to form TS1. Then, H*, O*, and CO* move to certain positions to form TS2. Next, one of the H* combines with O* to form OH*, while the other H* approaches OH* to form TS3, and then H* combines with OH* to form H2O*. Finally, H2O* and CO* desorb from CeO2−x, forming H2O(g) and CO(g).
As shown in Table 4, in the redox mechanism of the RWGS reaction catalyzed by CeO2−x, the total absorbed heat is greater than the total released heat, indicating that the reaction is endothermic. The reaction energy barriers required for the cleavage of the C=O bond, the formation of OH*, and the formation of H2O* are 1.86 eV, 1.43 eV, and 0.76 eV, respectively. In the redox mechanism of the RWGS reaction catalyzed by CeO2−x, the cleavage of the C=O bond requires the most energy, making the breaking of the C=O bond the rate-determining step in the redox mechanism of the CeO2−x catalyzed RWGS reaction.

2.5.2. Formate Association Mechanism of CeO2−x Catalyzed RWGS Reaction

As shown in Figure 6, in the formate-associated mechanism of the RWGS reaction catalyzed by CeO2−x, the initial adsorption states of CO2 and H2 are consistent with the redox pathway. CO2 adsorbs on CeO2−x, forming CO2*. H2 adsorbs on CeO2−x and dissociates into two H*. One of the H* approaches the C in CO2* to form TS1, H* combines with CO2* to form HCOO*, and then HCOO* is cleaved into HCO* and O* to form TS2. H*, O*, and HCO* move to certain positions to form TS3. Then HCO* is cleaved into CO* and H*. CO*, H*, and O* move to certain positions to form TS4. O* and H* form OH*, OH* and H* approach each other to form TS5, and OH* combines with H* to form H2O*. Finally, H2O* and CO* desorb from CeO2−x, forming H2O(g) and CO(g).
As shown in Table 5, in the formate-association mechanism of the RWGS reaction catalyzed by CeO2−x, the total heat absorbed is greater than the total heat released, making this an endothermic reaction. The reaction energy barriers required for the formation of HCOO*, HCO*, CO*, OH*, and H2O* are 0.72 eV, 1.93 eV, 1.58 eV, 1.42 eV, and 0.76 eV, respectively. In the formate-association mechanism for reverse water–gas shift catalyzed by CeO2−x, the energy required for the decomposition of HCOO* into HCO* is the highest. Therefore, the decomposition of HCOO* is the rate-determining step in the formate-associated mechanism of the RWGS reaction catalyzed by CeO2−x.

2.5.3. Carboxylate Association Mechanism of CeO2−x Catalyzed RWGS Reaction

As shown in Figure 7, in the carboxylate-associated mechanism of CeO2−x catalyzed RWGS reaction, the initial adsorption states of CO2 and H2 are consistent with the redox pathway. After CO2 and H2 molecules are adsorbed and activated on CeO2−x, H2 is dissociated into two H* atoms, one of which combines with CO2* to form trans-COOH*, after which trans-COOH* moves with H* to form TS2. Then one oxygen atom in trans-COOH* rotates by a certain angle, forming cis-COOH. cis-COOH* and H* move, forming TS3, then one O atom in cis-COOH* cleaves, forming COH* and O*. COH* rotates by a certain angle to form TS4, after which the H in COH* cleaves to form CO* and H*. At this moment, O* approaches one of the H* atoms to form TS5, O* combines with H* to form OH*, then OH* moves with another H* to form TS6, and OH* combines with H* to form H2O*. Finally, H2O* and CO* desorb from CeO2−x, forming H2O(g) and CO(g).
As shown in Table 6, in the carboxylate-associated mechanism of the CeO2−x catalyzed RWGS reaction, the total absorbed heat is greater than the total released heat, indicating that the reaction is endothermic. The reaction energy barriers required for the formation of trans-COOH*, cis-COOH*, COH*, CO*, OH*, and H2O* are 1.96 eV, 0.56 eV, 0.88 eV, 1.46 eV, 1.42 eV, and 0.66 eV, respectively. In the carboxylate-associated mechanism of reverse water–gas shift catalyzed by CeO2−x, the formation of COOH* requires the most energy. Therefore, the formation of COOH* is the rate-determining step in the carboxylate association mechanism of the reverse water–gas shift reaction catalyzed by CeO2−x.
As shown in Figure 8, the energies required for the rate-determining steps in the redox pathway, formate-associated pathway, and carboxylate-associated mechanism of CeO2−x catalyzed RWGS reaction are 1.86 eV, 1.93 eV, and 1.96 eV, respectively. The rate-determining step of the redox mechanism requires the least energy, so the most favorable reaction mechanism for CeO2−x catalyzed RWGS reaction is the redox pathway, indicating that CeO2−x catalyzed RWGS reaction follows a redox mechanism. This is consistent with the findings of Liu et al. [35] that the direct oxidative decomposition of carbon dioxide at oxygen vacancies on CeO2 catalysts is considered the main reaction mechanism of the RWGS.

2.5.4. Redox Mechanisms of the Ni/CeO2 Catalyzed RWGS Reaction

As shown in Figure 9, in the redox mechanism of the RWGS reaction catalyzed by Ni/CeO2, CO2 adsorbs on Ni/CeO2 and undergoes significant distortion, indicating that CO2* is activated. H2 molecules adsorbed onto Ni atoms are directly cleaved into two H*, indicating the excellent H2 dissociation and activation capability of Ni-based catalysts. Then CO2* decomposes into CO* and O*, forming TS1.O* and H* move positions to form TS2. Then, O* combines with H* to form OH*. OH* and H* move a certain distance to form TS3. Then OH* combines with H* to form H2O*. Finally, CO* and H2O* absorb a certain amount of heat, desorb from Ni/CeO2, and form CO(g) and H2O(g).
As shown in Table 7, in the redox mechanism of the Ni/CeO2 catalyzed RWGS reaction, the total endothermic energy is greater than the total exothermic energy, indicating that the reaction is endothermic. The energy required to break the C=O bond is 1.78 eV, which is the highest energy needed; therefore, the breaking of the C=O bond is the rate-determining step in the redox mechanism of the Ni/CeO2 catalyzed RWGS reaction.

2.5.5. Formate Association Mechanism of Ni/CeO2 Catalyzed RWGS Reaction

As shown in Figure 10, the initial adsorption states of CO2 and H2 in the formate association mechanism are consistent with those in the redox pathway. CO2 adsorbed on Ni/CeO2 undergoes deformation, and H2 molecules adsorbed on Ni atoms are directly dissociated into two H*. H* approaches the carbon atom in CO2* to form TS1. Next, H* combines with CO2* to form HCOO*. HCOO* absorbs a certain amount of energy, causing the C=O bond to break into HCO* and O*, forming TS2. HCO* twists at a certain angle to form TS3. Next, the C-H bond in HCO* breaks, forming CO* and H*. O* and H* move a certain distance to form TS4. When the distance between O* and one of the H* atoms is sufficiently close, O* combines with H* to form OH*, which twists at a certain angle to form TS5. Then OH* combines with another H* to form H2O*. Finally, CO* and H2O* absorb a certain amount of heat, desorb from Ni/CeO2, and form CO(g) and H2O(g).
As shown in Table 8, in the formate-associated mechanism of the Ni/CeO2 catalyzed RWGS reaction, the absorbed heat is greater than the released heat, making this an endothermic reaction. The reaction energy barriers required for the formation of HCOO*, HCO*, CO*, OH*, and H2O* are 0.63 eV, 1.57 eV, 1.29 eV, 1.37 eV, and 0.62 eV, respectively. In the Ni/CeO2 catalyzed RWGS reaction via the formate-associated pathway, the energy required for HCOO* to decompose into HCO* is the highest, making the decomposition of HCOO* the rate-determining step in the formate-associated mechanism of the Ni/CeO2 catalyzed RWGS reaction.

2.5.6. Carboxylate Association Mechanism of Ni/CeO2 Catalyzed RWGS Reaction

As shown in Figure 11, after the CO2 and H2 molecules are adsorbed and activated, CO2* approaches one of the H*, forming TS1. CO2* combines with H* to form trans-COOH*, and one O in trans-COOH* twists by a certain angle to form TS2. When the O in trans-COOH* twists to a certain angle, it forms cis-COOH*. Then, the C=O bond in cis-COOH elongates, forming TS3. When the C=O bond in cis-COOH is stretched to a certain distance, it breaks, forming COH* and O*. Subsequently, COH* is cleaved into CO* and H*, forming TS4. O* approaches one of the H* atoms, forming TS5. When the distance between O* and H* is close enough, OH* is formed. OH* twists at a certain angle to approach another H*, forming TS6. OH* combines with H* to form H2O*. Finally, CO* and H2O* absorb a certain amount of heat and desorb from Ni/CeO2, forming CO(g) and H2O(g).
As shown in Table 9, in the carboxylate associative mechanism of the RWGS reaction catalyzed by Ni/CeO2, the absorbed heat is greater than the released heat, making the reaction endothermic. The reaction energy barrier required for the formation of trans-COOH* is 1.85 eV, which is the highest energy needed in the carboxylate association mechanism of the RWGS reaction catalyzed by Ni/CeO2. Therefore, the formation of COOH* is the rate-determining step in the carboxylate association mechanism of the RWGS reaction catalyzed by Ni/CeO2.
As can be seen from Figure 12, the energy required for the rate-determining steps of the redox mechanism, formate-associated mechanism, and carboxylate-associated mechanism of Ni/CeO2 catalyzed RWGS reaction are 1.78 eV, 1.57 eV, and 1.85 eV, respectively. In the Ni/CeO2 catalyzed RWGS reaction, the rate-determining step along the formate mechanism requires the least energy. Therefore, the most favorable reaction mechanism in the Ni/CeO2 catalyzed RWGS reaction is the formate mechanism, indicating that single-atom Ni-loaded CeO2 catalyzes the RWGS reaction following the formate-association mechanism. This is consistent with Yu et al. [42], who found that CO mainly originates from the decomposition of formate species.

3. Calculation Details

All density functional calculations were performed using the DMol3 module of Materials Studio [43]. The exchange-correlation function used the generalized gradient approximation (GGA) and the Perdew-Burke-Ernzerhof (PBE) equation [44]. The core is treated with DFT semi-core pseudopotentials (DSPP) for self-consistent field calculations. The atomic orbital basis set is based on the double numerical plus polarization (DNP) basis set. The integrations of the Brillouin zone were described using a 2 × 2 × 1 k-point grid according to the Monkhorst-Pack method. The convergence tolerance of geometric optimization and property calculation were 1 × 10−5 Ha for the tolerance of energy, 0.002 Ha/Å for maximum force, 0.005 Å for maximum displacement and 1.0 × 10−6 Ha for self-consistent field (SCF) tolerance. Studies have shown that the PBE functional can be used to obtain accurate structural information in calculations involving transition metals and transition metal oxides [45,46]. The transition state of the reaction was determined using the full LST/QST method [47].
The definition of adsorption energy in this article is shown in Equation (1):
Eads = Eadsorbate-catalyst − Eadsorbate − Ecatalyst
Eadsorbate-catalyst represents the total energy of the adsorbate and the catalyst when the adsorbate is stable on the catalyst surface, while Eadsorbate and Ecatalyst represent the total energy of the adsorbate and the catalyst, respectively.
The energy required to move from reactant to transition state on a catalyst in this paper is shown in Equation (2):
ΔE = ETS − ER
where ΔE represents the energy required for the reactants to the transition state, and ER represents the energy of the reactants adsorbed on the catalyst. ETS represents the energy of the transition state adsorbed on the catalyst.
All catalyst models use a periodic 3 × 3 supercell representation with six layers and 20 Å of vacuum to avoid interactions between periodic images. The bottom four atomic layers are fixed, while the top two layers and the adsorbates are fully relaxed without any constraints. The CeO2(111) surface is the first stable surface of CeO2, while the CeO2(110) surface is its second stable surface. Therefore, we chose the CeO2(111) surface for our study. An oxygen atom is removed from the outermost CeO2 layer to form an oxygen vacancy CeO2−x(111). A Ni atom is loaded onto the CeO2, forming Ni/CeO2. The front and top views of the structural models of CeO2, CeO2−x, and Ni/CeO2 are shown in Figure 13 and Figure 14.

4. Conclusions

This study employed DFT to systematically investigate the RWGS reaction over three CeO2-based catalysts: pure CeO2, CeO2−x, and Ni/CeO2, with key findings summarized as follows: First, regarding CO2 adsorption sites: DFT calculations revealed that CO2 is more easily adsorbed at the intermediate oxygen sites of CeO2 and Ni/CeO2, while it preferentially binds to the oxygen vacancies of CeO2−x. Second, on catalytic activity: Through Mulliken charge analysis and band gap calculations, the catalytic activities of the three catalysts were compared. The results indicated that CeO2−x and Ni/CeO2 exhibit significantly higher activity than pure CeO2. For Ni/CeO2, the enhanced activity is closely related to the electronic interaction between Ni and the CeO2 support, which modulates the electronic structure of active sites and promotes CO2 activation. Third, regarding the CO2 activation mechanism: the activation mechanisms of different CeO2-based catalysts were revealed using density of states analysis. The study found that all three catalysts can activate CO2 to varying degrees; notably, there is strong hybridization between the d orbitals of Ni and the p orbitals of oxygen in CO2 in Ni/CeO2, which serves as the core electronic driving force for the high activation efficiency of Ni/CeO2. Fourth, regarding the RWGS reaction mechanism and rate-determining steps: the study further explored the reaction mechanisms of CeO2−x and Ni/CeO2 in catalytic RWGS reactions, as well as the rate-determining steps of each mechanism. The results show that for RWGS reactions catalyzed by these two catalysts: the rate-determining step of the redox mechanism is the breaking of the C=O bond; the rate-determining step of the formate mechanism is the decomposition of HCOO*; and the rate-determining step of the carboxylate mechanism is the formation of COOH*. Finally, regarding the dominant mechanism of the catalyst: key studies on the mechanisms of CeO2−x and Ni/CeO2 confirmed that CeO2−x catalyzes the RWGS reaction through the redox mechanism, while Ni/CeO2 follows a mechanism related to formate. This clear understanding of the dominant mechanisms provides a precise theoretical basis for the design and optimization of CeO2−x and Ni/CeO2 catalysts in RWGS applications.
Although this study conducted an in-depth investigation of the RWGS reaction on CeO2, CeO2−x, and Ni/CeO2 using DFT calculations, certain challenges still exist. Traditional DFT methods have obvious limitations when dealing with strongly correlated systems, and they cannot fully account for defect sites commonly present in actual catalytic systems (such as multiple oxygen vacancies and surface hydroxyl groups) as well as dynamic structural evolution processes (such as catalyst surface reconstruction). These limitations may lead to discrepancies between theoretical predictions and experimental results. To address these issues, it is recommended that future research focus on the following directions: constructing a closed-loop research paradigm of computational design–machine learning optimization–experimental validation, integrating theoretical predictions with experimental verification to improve the reliability of catalyst design, and establishing standardized multi-source databases, integrating experimental data (such as catalytic activity and structural characterization) with simulation data (such as DFT-calculated adsorption energies and electronic structures), to facilitate data sharing and cross-validation.

Author Contributions

The contributions of the authors for the manuscript are the following: conceptualization, W.X. and D.L.; methodology, W.X. and K.C.; software, X.W. and B.L.; validation, D.W.; formal analysis, D.W. and X.W.; investigation, B.L.; resources, B.L. and M.D.; data curation, Y.Z. and X.W.; writing—original draft, X.W., Y.Z. and B.L.; writing—review and editing, K.C. and D.L.; visualization, M.D.; supervision, W.X.; project administration, W.X.; funding acquisition, W.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work is a project sponsored by the National Natural Science Foundation of China (Grant 21978327), Shandong Provincial Natural Science Foundation, China (Grant ZR2023MB005).

Data Availability Statement

The data presented in this study are available within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, W.; Wang, S.; Ma, X. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703–3727. [Google Scholar] [CrossRef]
  2. Pan, S.Y.; Chang, E.E.; Chiang, P.C. CO2 capture by accelerated carbonation of alkaline wastes: A review on its principles and applications. Aerosol Air Qual. Res. 2012, 12, 770–791. [Google Scholar] [CrossRef]
  3. Li, L.; Zhao, N.; Wei, W. A review of research progress on CO2 capture, storage, and utilization in Chinese Academy of Sciences. Fuel 2013, 108, 112–130. [Google Scholar] [CrossRef]
  4. Artz, J.; Muller, T.E.; Thenert, K. Sustainable conversion of carbon dioxide: An integrated review of catalysis and life cycle assessment. Chem. Rev. 2018, 118, 434–504. [Google Scholar] [CrossRef]
  5. Yang, H.; Zhang, C.; Gao, P. A review of the catalytic hydrogenation of carbon dioxide into value-added hydrocarbons. Catal. Sci. Technol. 2017, 7, 4580–4598. [Google Scholar] [CrossRef]
  6. Zhang, C.; Li, P.; Xiao, Z. Tuning the CO2 hydrogenation activity and selectivity by highly dispersed Ni–In intermetallic alloy compounds. ACS Sustain. Chem. Eng. 2023, 12, 166–177. [Google Scholar] [CrossRef]
  7. Liu, Y.; Li, L.; Zhang, R. Synergetic enhancement of activity and selectivity for reverse water gas shift reaction on Pt-Re/SiO2 catalysts. J. CO2 Util. 2022, 63, 102128. [Google Scholar] [CrossRef]
  8. Ou, Z.; Qin, C.; Niu, J. A comprehensive DFT study of CO2 catalytic conversion by H2 over Pt-doped Ni catalysts. Int. J. Hydrogen Energy 2019, 44, 819–834. [Google Scholar] [CrossRef]
  9. Kim, S.S.; Lee, H.H.; Hong, S.C. A study on the effect of support’s reducibility on the reverse water-gas shift reaction over Pt catalysts. Appl. Catal. A Gen. 2012, 423, 100–107. [Google Scholar] [CrossRef]
  10. Wang, K.; Cao, M.; Lu, J. Operando DRIFTS-MS investigation on plasmon-thermal coupling mechanism of CO2 hydrogenation on Au/TiO2: The enhanced generation of oxygen vacancies. Appl. Catal. B Environ. 2021, 296, 120341. [Google Scholar] [CrossRef]
  11. Pettigrew, D.J.; Trimm, D.L.; Cant, N.W. The effects of rare earth oxides on the reverse water-gas shift reaction on palladium/alumina. Catal. Lett. 1994, 28, 313–319. [Google Scholar] [CrossRef]
  12. Zhou, G.; Xie, F.; Deng, L. Supported mesoporous Cu/CeO2−δ catalyst for CO2 reverse water–gas shift reaction to syngas. Int. J. Hydrogen Energy 2020, 45, 11380–11393. [Google Scholar] [CrossRef]
  13. Peng, G.; Sibener, S.J.; Schatz, G.C. CO2 hydrogenation to formic acid on Ni(111). J. Phys. Chem. C 2012, 116, 3001–3006. [Google Scholar] [CrossRef]
  14. Rodriguez, J.A.; Evans, J.; Feria, L. CO2 hydrogenation on Au/TiC, Cu/TiC, and Ni/TiC catalysts: Production of CO, methanol, and methane. J. Catal. 2013, 307, 162–169. [Google Scholar] [CrossRef]
  15. Fu, B.; Zhao, Y.; Zhang, S. Effectively regulating Ni distribution on CeO2 through formation of Ce1-xNixO2-y solid solution to enhance the selectivity in RWGS reaction. Fuel 2025, 382, 133669. [Google Scholar] [CrossRef]
  16. Trovarelli, A. Catalytic properties of ceria and CeO2-containing materials. Catal. Rev. 1996, 38, 439–520. [Google Scholar] [CrossRef]
  17. Zou, J.; Oladipo, J.; Fu, S. Hydrogen production from cellulose catalytic gasification on CeO2/Fe2O3 catalyst. Energy Convers. Manag. 2018, 171, 241–248. [Google Scholar] [CrossRef]
  18. Kim, S.S.; Lee, H.H.; Hong, S.C. The effect of the morphological characteristics of TiO2 supports on the reverse water–gas shift reaction over Pt/TiO2 catalysts. Appl. Catal. B Environ. 2012, 119, 100–108. [Google Scholar] [CrossRef]
  19. Wang, X.; Shi, H.; Kwak, J.H. Mechanism of CO2 hydrogenation on Pd/Al2O3 catalysts: Kinetics and transient DRIFTS-MS studies. ACS Catal. 2015, 5, 6337–6349. [Google Scholar] [CrossRef]
  20. Wang, L.; Zhang, S.; Liu, Y. Reverse water gas shift reaction over Co-precipitated Ni-CeO2 catalysts. J. Rare Earths 2008, 26, 66–70. [Google Scholar] [CrossRef]
  21. Chen, X.; Su, X.; Duan, H. Catalytic performance of the Pt/TiO2 catalysts in reverse water gas shift reaction: Controlled product selectivity and a mechanism study. Catal. Today 2017, 281, 312–318. [Google Scholar] [CrossRef]
  22. Lu, B.; Kawamoto, K. Preparation of mesoporous CeO2 and monodispersed NiO particles in CeO2, and enhanced selectivity of NiO/CeO2 for reverse water gas shift reaction. Mater. Res. Bull. 2014, 53, 70–78. [Google Scholar] [CrossRef]
  23. Chen, C.S.; Budi, C.S.; Wu, H.C. Size-tunable Ni nanoparticles supported on surface-modified, cage-type mesoporous silica as highly active catalysts for CO2 hydrogenation. ACS Catal. 2017, 7, 8367–8381. [Google Scholar] [CrossRef]
  24. Zhang, X.; Han, S.; Zhu, B. Reversible loss of core–shell structure for Ni–Au bimetallic nanoparticles during CO2 hydrogenation. Nat. Catal. 2020, 3, 411–417. [Google Scholar] [CrossRef]
  25. Yang, X.; Su, X.; Chen, X. Promotion effects of potassium on the activity and selectivity of Pt/zeolite catalysts for reverse water gas shift reaction. Appl. Catal. B Environ. 2017, 216, 95–105. [Google Scholar] [CrossRef]
  26. Liang, B.; Duan, H.; Su, X. Promoting role of potassium in the reverse water gas shift reaction on Pt/mullite catalyst. Catal. Today 2017, 281, 319–326. [Google Scholar] [CrossRef]
  27. Wu, H.C.; Chang, Y.C.; Wu, J.H. Methanation of CO2 and reverse water gas shift reactions on Ni/SiO2 catalysts: The influence of particle size on selectivity and reaction pathway. Catal. Sci. Technol. 2015, 5, 4154–4163. [Google Scholar] [CrossRef]
  28. Xiao, Z.; Zhang, C.; Gu, J. Designing Ni-In intermetallic alloy compounds for high activity and selectivity in low-temperature RWGS reaction. Chem. Eng. J. 2025, 507, 160529. [Google Scholar] [CrossRef]
  29. Matsubu, J.C.; Yang, V.N.; Christopher, P. Isolated metal active site concentration and stability control catalytic CO2 reduction selectivity. J. Am. Chem. Soc. 2015, 137, 3076–3084. [Google Scholar] [CrossRef]
  30. Zhang, J.; Mao, Y.; Zhang, J. CO2 reforming of ethanol: Density functional theory calculations, microkinetic modeling, and experimental studies. ACS Catal. 2020, 10, 9624–9633. [Google Scholar] [CrossRef]
  31. Hu, X.; Tao, L.; Lei, K.; Sun, Z.; Tan, M. Unraveling TiO2 phase effects on Pt single-atom catalysts for efficient CO2 conversion. Chin. J. Catal. 2025, 73, 186–195. [Google Scholar] [CrossRef]
  32. Yang, B.; Wang, Y.; Gao, B. Size-dependent active site and its catalytic mechanism for CO2 hydrogenation reactivity and selectivity over Re/TiO2. ACS Catal. 2023, 13, 10364–10374. [Google Scholar] [CrossRef]
  33. Li, Y.; Zhao, Z.; Lu, W. Single-atom Co-N-C catalysts for high-efficiency reverse water-gas shift reaction. Appl. Catal. B Environ. 2023, 324, 122298. [Google Scholar] [CrossRef]
  34. Zhou, G.; Dai, B.; Xie, H. CeCu composite catalyst for CO synthesis by reverse water–gas shift reaction: Effect of Ce/Cu mole ratio. J. CO2 Util. 2017, 21, 292–301. [Google Scholar] [CrossRef]
  35. Liu, Y.; Li, Z.; Xu, H. Reverse water–gas shift reaction over ceria nanocube synthesized by hydrothermal method. Catal. Commun. 2016, 76, 1–6. [Google Scholar] [CrossRef]
  36. Rabee, A.I.M.; Abed, H.; Vuong, T.H. CeO2-supported single-atom Cu catalysts modified with Fe for RWGS reaction: Deciphering the role of Fe in the reaction mechanism by in situ/operando spectroscopic techniques. ACS Catal. 2024, 14, 10913–10927. [Google Scholar] [CrossRef]
  37. Choi, Y.; Sim, G.D.; Jung, U. Copper catalysts for CO2 hydrogenation to CO through reverse water–gas shift reaction for e-fuel production: Fundamentals, recent advances, and prospects. Chem. Eng. J. 2024, 492, 152283. [Google Scholar] [CrossRef]
  38. Jangam, A.; Das, S.; Dewangan, N. Conversion of CO2 to C1 chemicals: Catalyst design, kinetics and mechanism aspects of the reactions. Catal. Today 2020, 358, 3–29. [Google Scholar] [CrossRef]
  39. Su, X.; Yang, X.; Zhao, B. Designing of highly selective and high-temperature endurable RWGS heterogeneous catalysts: Recent advances and the future directions. J. Energy Chem. 2017, 26, 854–867. [Google Scholar] [CrossRef]
  40. Wang, K.; Shao, S.; Liu, Y. DRIFTS-SSITKA-MS investigations on the mechanism of plasmon preferentially enhanced CO2 hydrogenation over Au/γ-Al2O3. Appl. Catal. B Environ. 2023, 328, 122531. [Google Scholar] [CrossRef]
  41. Yao, X.; Wei, Z.; Mei, J. The reverse water gas shift reaction (RWGS) mechanism study on the γ-MoC(100) surface. RSC Adv. 2025, 15, 460–466. [Google Scholar] [CrossRef]
  42. Yu, Y.; Bian, Z.; Wang, Z. CO2 methanation on Ni-Ce0.8M0.2O2 (M=Zr, Sn or Ti) catalyst: Suppression of CO via formation of bridging carbonyls on nickel. Catal. Today 2023, 424, 113053. [Google Scholar] [CrossRef]
  43. Delley, B. From molecules to solids with the DMol3 approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  44. Liu, R.; Zhang, C.; Chu, W. Enhancing catalytic efficiency in zeolite-supported Pt nanoclusters for RWGS reaction through potassium incorporation: A DFT study. Int. J. Hydrogen Energy 2025, 104, 131–137. [Google Scholar] [CrossRef]
  45. Zhang, M.; Huang, Y.; Li, R. A DFT Study of Ethanol Adsorption and Dehydrogenation on Cu/Cr2O3 Catalyst. Catal. Lett. 2014, 144, 1978–1986. [Google Scholar] [CrossRef]
  46. Miao, B.; Wu, Z.P.; Xu, H. DFT studies on the key competing reaction steps towards complete ethanol oxidation on transition metal catalysts. Comput. Mater. Sci. 2019, 156, 175–186. [Google Scholar] [CrossRef]
  47. Zhang, M.; Zhuang, J.; Yu, Y. A DFT study on ZrO2 surface in the process of ethanol to 1,3-butadiene: A comprehensive mechanism elucidation. Appl. Surf. Sci. 2018, 458, 1026–1034. [Google Scholar] [CrossRef]
Figure 1. The top two layers of Ce atoms and O atoms in different CeO2. Note: The red ball, yellow ball, and blue ball represent oxygen, cerium, and nickel atoms, respectively.
Figure 1. The top two layers of Ce atoms and O atoms in different CeO2. Note: The red ball, yellow ball, and blue ball represent oxygen, cerium, and nickel atoms, respectively.
Catalysts 15 01028 g001
Figure 2. Stable configurations of CO2 adsorption on different CeO2 surfaces. 1: CeO2, 2: CeO2−x, 3: Ni/CeO2. Note: The red, yellow, blue, and gray balls represent oxygen, cerium, nickel, and carbon atoms, respectively. Dashed line: the distance from the carbon atom in CO2 to different outermost oxygen atom sites or nickel atom sites in different CeO2.
Figure 2. Stable configurations of CO2 adsorption on different CeO2 surfaces. 1: CeO2, 2: CeO2−x, 3: Ni/CeO2. Note: The red, yellow, blue, and gray balls represent oxygen, cerium, nickel, and carbon atoms, respectively. Dashed line: the distance from the carbon atom in CO2 to different outermost oxygen atom sites or nickel atom sites in different CeO2.
Catalysts 15 01028 g002
Figure 3. (ac) Band gaps of different CeO2(111); (df) band gaps of different CeO2 after CO2 adsorption.
Figure 3. (ac) Band gaps of different CeO2(111); (df) band gaps of different CeO2 after CO2 adsorption.
Catalysts 15 01028 g003
Figure 4. Density of states diagrams of CO2 adsorbed on CeO2 (ac), CeO2−x (df), and Ni/CeO2 (gi).
Figure 4. Density of states diagrams of CO2 adsorbed on CeO2 (ac), CeO2−x (df), and Ni/CeO2 (gi).
Catalysts 15 01028 g004
Figure 5. Intermediates and transition states in the redox mechanism of CeO2−x catalyzed RWGS reaction. *: adsorbed species. Note: Red balls, white balls, gray balls, and yellow balls represent oxygen atoms, hydrogen atoms, carbon atoms, and cerium atoms, respectively.
Figure 5. Intermediates and transition states in the redox mechanism of CeO2−x catalyzed RWGS reaction. *: adsorbed species. Note: Red balls, white balls, gray balls, and yellow balls represent oxygen atoms, hydrogen atoms, carbon atoms, and cerium atoms, respectively.
Catalysts 15 01028 g005
Figure 6. Intermediates and transition states in the formate-association mechanism for CeO2−x catalyzed RWGS reaction. *: adsorbed species.
Figure 6. Intermediates and transition states in the formate-association mechanism for CeO2−x catalyzed RWGS reaction. *: adsorbed species.
Catalysts 15 01028 g006
Figure 7. Intermediates and transition states in the carboxylate association mechanism of CeO2−x catalyzed RWGS reaction. *: adsorbed species.
Figure 7. Intermediates and transition states in the carboxylate association mechanism of CeO2−x catalyzed RWGS reaction. *: adsorbed species.
Catalysts 15 01028 g007
Figure 8. Energy distribution diagram of three mechanisms (redox, formate, and carboxylate mechanisms) for catalyzing the RWGS reaction on CeO2−x. *: adsorbed species.
Figure 8. Energy distribution diagram of three mechanisms (redox, formate, and carboxylate mechanisms) for catalyzing the RWGS reaction on CeO2−x. *: adsorbed species.
Catalysts 15 01028 g008
Figure 9. Intermediates and transition states in the redox mechanism of the Ni/CeO2 catalyzed RWGS reaction. *: adsorbed species.
Figure 9. Intermediates and transition states in the redox mechanism of the Ni/CeO2 catalyzed RWGS reaction. *: adsorbed species.
Catalysts 15 01028 g009
Figure 10. Intermediates and transition states in the formate association mechanism for the Ni/CeO2 catalyzed RWGS reaction. *: adsorbed species.
Figure 10. Intermediates and transition states in the formate association mechanism for the Ni/CeO2 catalyzed RWGS reaction. *: adsorbed species.
Catalysts 15 01028 g010
Figure 11. Intermediates and transition states in the carboxylate-associated mechanism of the Ni/CeO2 catalyzed RWGS reaction. *: adsorbed species.
Figure 11. Intermediates and transition states in the carboxylate-associated mechanism of the Ni/CeO2 catalyzed RWGS reaction. *: adsorbed species.
Catalysts 15 01028 g011
Figure 12. Energy distribution diagram of the three mechanisms (redox, formate, and carboxylate mechanisms) for catalyzing the RWGS reaction on Ni/CeO2. *: adsorbed species.
Figure 12. Energy distribution diagram of the three mechanisms (redox, formate, and carboxylate mechanisms) for catalyzing the RWGS reaction on Ni/CeO2. *: adsorbed species.
Catalysts 15 01028 g012
Figure 13. Front view of the structural models of CeO2, CeO2−x, and Ni/CeO2. (Red, yellow, and blue represent oxygen, cerium, and nickel atoms, respectively).
Figure 13. Front view of the structural models of CeO2, CeO2−x, and Ni/CeO2. (Red, yellow, and blue represent oxygen, cerium, and nickel atoms, respectively).
Catalysts 15 01028 g013
Figure 14. Top view of the structural models of CeO2, CeO2−x, and Ni/CeO2.
Figure 14. Top view of the structural models of CeO2, CeO2−x, and Ni/CeO2.
Catalysts 15 01028 g014
Table 1. Mulliken charges of Ce atoms and O atoms.
Table 1. Mulliken charges of Ce atoms and O atoms.
AtomCeO2(111)CeO2−x(111)Ni/CeO2(111)
Ce11.5991.7461.600
Ce21.5991.5451.786
Ce31.5991.5431.574
O1−0.799−0.859−0.801
O2−0.799-−0.809
O3−0.799−0.859−0.810
Table 2. Adsorption energies and geometric parameters of CO2 on different CeO2 surfaces.
Table 2. Adsorption energies and geometric parameters of CO2 on different CeO2 surfaces.
ConfigurationEads (eV)C=O (Å)InteractionDistance (Å)
1A−0.7891.198, 1.232C-O11.405
1B−0.7761.186, 1.227C-O21.408
2O-vacancy−1.1701.233, 1.249C-Ovacancy-
2B−1.0241.225, 1.241C-O21.403
3A−1.2521.257, 1.266C-O11.382
3B−1.2321.248, 1.256C-O21.402
3C−1.2361.251, 1.259C-Ni1.810
Note: O1 and O2 represent intermediate oxygen and edge oxygen of CeO2-based catalysts, respectively.
Table 3. Basic steps involved in the RWGS mechanism.
Table 3. Basic steps involved in the RWGS mechanism.
Redox Mechanism [37,38]Formate Mechanism [39]Carboxylate Mechanism [40,41]
H2 + 2 * → 2H*H2(g) + 2 * → 2H*H2(g) + 2 * → 2H*
CO2(g) + * → CO2*CO2(g) + * → CO2*CO2(g) + * → CO2*
CO2* + * → CO* + O*H* + CO2* → HCOO* + *CO2* + H* → trans-COOH* + H*
O* + H* → OH*HCOO* → HCO* + O*trans-COOH* → cis-COOH*
OH* + H* → H2O* + *HCO* → CO* + H*cis-COOH* → COH* + O*
H2O* → * + H2O(g)H* + O* → OH* + *COH* + * → CO* + H*
CO* → * + CO(g)H* + OH* → H2O* + *H* + O* → OH* + *
H2O* → *+ H2O(g)H* + OH* → H2O* + *
CO* → *+ CO(g)H2O* → * + H2O(g)
CO* → * + CO(g)
Note: * indicates species adsorbed on the catalyst surface.
Table 4. Reaction energy barriers and reaction heats in the redox process of CeO2−x catalyzed RWGS.
Table 4. Reaction energy barriers and reaction heats in the redox process of CeO2−x catalyzed RWGS.
Redox PathwayReaction Energy Barrier ΔE/eVReaction Heat ΔH/eV
CO2(g) + H2(g) → CO2* + H2(g)-−1.12
CO2* + H2(g) → CO2* + 2H*-−0.32
CO2* + 2H* → CO* + O* + 2H*1.86−0.28
CO* + O* + 2H* → CO* + OH* + H*1.430.35
CO* + OH* + H* → CO* + H2O*0.760.46
CO* + H2O* → CO* + H2O(g)-0.81
CO* + H2O(g) → CO(g) + H2O(g)-1.08
Note: * indicates species adsorbed on the catalyst surface.
Table 5. Reaction energy barriers and reaction enthalpies in the formate-mechanism of the RWGS reaction catalyzed by CeO2−x.
Table 5. Reaction energy barriers and reaction enthalpies in the formate-mechanism of the RWGS reaction catalyzed by CeO2−x.
Formate PathwayReaction Energy Barrier ΔE/eVReaction Heat ΔH/eV
CO2(g) + H2(g) → CO2* + H2(g)-−1.12
CO2* + H2(g) → CO2* + 2H*-−0.32
CO2* + 2H* → HCOO* + H*0.72−0.09
HCOO* + H* → HCO* + O* + H*1.93−0.18
HCO* + O* + H* → CO* + O* + 2H*1.58−0.16
CO* + O* + 2H* → CO* + OH* + H*1.420.35
CO* + OH* + H* → CO* + H2O*0.760.46
CO* + H2O* → CO* + H2O(g)-0.81
CO* + H2O(g) → CO(g) + H2O(g)-1.08
Note: * indicates species adsorbed on the catalyst surface.
Table 6. Reaction energy barriers and reaction enthalpies in the carboxylate association mechanism of CeO2−x catalyzed RWGS reaction.
Table 6. Reaction energy barriers and reaction enthalpies in the carboxylate association mechanism of CeO2−x catalyzed RWGS reaction.
Carboxylate PathwayReaction Energy Barrier ΔE/eVReaction Heat ΔH/eV
CO2(g) + H2(g) → CO2* + H2(g)-−1.12
CO2* + H2(g) → CO2* + 2H*-−0.32
CO2* + 2H* → trans-COOH* + H*1.960.52
trans-COOH* + H* → cis-COOH* + H*0.560.12
cis-COOH* + H* → COH* + O* + H*0.88−0.35
COH* + O* + H* → CO* + O* + 2H*1.46−0.42
CO* + O* + 2H* → CO* + OH* + H*1.420.35
CO* + OH* + H* → CO* + H2O*0.660.42
CO* + H2O* → CO* + H2O(g)-0.81
CO* + H2O(g) → CO(g) + H2O(g)-1.08
Note: * indicates species adsorbed on the catalyst surface.
Table 7. Reaction energy barriers and reaction heats in the Ni/CeO2 catalyzed RWGS redox process.
Table 7. Reaction energy barriers and reaction heats in the Ni/CeO2 catalyzed RWGS redox process.
Redox PathwayReaction Energy Barrier ΔE/eVReaction Heat ΔH/eV
CO2(g) + H2(g) → CO2* + H2(g)-−1.17
CO2* + H2(g) → CO2* + 2H*-−0.36
CO2* + 2H* → CO*+O* + 2H*1.78−0.26
CO* + O* + 2H* → CO* + OH* + H*1.370.34
CO* + OH* + H* → CO* + H2O*0.620.41
CO* + H2O* → CO* + H2O(g)-0.76
CO* + H2O(g) → CO(g) + H2O(g)-1.03
Note: * indicates species adsorbed on the catalyst surface.
Table 8. Reaction energy barriers and reaction heats in the formate-associated mechanism of the Ni/CeO2 catalyzed RWGS reaction.
Table 8. Reaction energy barriers and reaction heats in the formate-associated mechanism of the Ni/CeO2 catalyzed RWGS reaction.
Formate PathwayReaction Energy Barrier ΔE/eVReaction Heat ΔH/eV
CO2(g) + H2(g) → CO2* + H2(g)-−1.17
CO2* + H2(g) → CO2* + 2H*-−0.36
CO2* + 2H* → HCOO* + H*0.63−0.11
HCOO* + H* → HCO* + O* + H*1.57−0.21
HCO* + O* + H* → CO* + O* + 2H*1.29−0.18
CO* + O* + 2H* → CO* + OH* +H*1.370.34
CO* + OH* + H* → CO* + H2O*0.620.41
CO* + H2O* → CO* + H2O(g)-0.76
CO* + H2O(g) → CO(g) + H2O(g)-1.03
Note: * indicates species adsorbed on the catalyst surface.
Table 9. Carboxylate Association Mechanisms of Ni/CeO2 Catalyzed RWGS Reaction.
Table 9. Carboxylate Association Mechanisms of Ni/CeO2 Catalyzed RWGS Reaction.
Carboxylate PathwayReaction Energy Barrier ΔE/eVReaction Heat ΔH/eV
CO2(g) + H2(g) ⟶ CO2* + H2(g)-−1.17
CO2* + H2(g) ⟶ CO2* + 2H*-−0.36
CO2* + 2H* ⟶ trans-COOH* + H*1.850.56
trans-COOH* + H* ⟶ cis-COOH* + H*0.420.08
cis-COOH* + H* ⟶ COH* + O* + H*0.76−0.36
COH* + O* + H* ⟶ CO* + O* + 2H*1.39−0.43
CO* + O* + 2H* ⟶ CO* + OH* + H*1.370.34
CO* + OH* + H* ⟶ CO* + H2O*0.620.41
CO* + H2O* ⟶ CO* + H2O(g)-0.76
CO* + H2O(g) ⟶ CO(g) + H2O(g)-1.03
Note: * indicates species adsorbed on the catalyst surface.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, X.; Xia, W.; Zhang, Y.; Wang, D.; Dong, M.; Chen, K.; Liu, D.; Lu, B. Unraveling the Reaction Mechanism of the Reverse Water–Gas Shift Reaction over Ni/CeO2 and CeO2−x Catalysts. Catalysts 2025, 15, 1028. https://doi.org/10.3390/catal15111028

AMA Style

Wang X, Xia W, Zhang Y, Wang D, Dong M, Chen K, Liu D, Lu B. Unraveling the Reaction Mechanism of the Reverse Water–Gas Shift Reaction over Ni/CeO2 and CeO2−x Catalysts. Catalysts. 2025; 15(11):1028. https://doi.org/10.3390/catal15111028

Chicago/Turabian Style

Wang, Xinrui, Wei Xia, Yanli Zhang, Di Wang, Mingyuan Dong, Kun Chen, Dong Liu, and Baowang Lu. 2025. "Unraveling the Reaction Mechanism of the Reverse Water–Gas Shift Reaction over Ni/CeO2 and CeO2−x Catalysts" Catalysts 15, no. 11: 1028. https://doi.org/10.3390/catal15111028

APA Style

Wang, X., Xia, W., Zhang, Y., Wang, D., Dong, M., Chen, K., Liu, D., & Lu, B. (2025). Unraveling the Reaction Mechanism of the Reverse Water–Gas Shift Reaction over Ni/CeO2 and CeO2−x Catalysts. Catalysts, 15(11), 1028. https://doi.org/10.3390/catal15111028

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop