Abstract
The reverse water–gas shift (RWGS) reaction is a key process for CO2 conversion and sustainable fuel production, yet the nature of the active sites on Pt/TiO2 cluster catalysts remains elusive. Using first-principles microkinetic simulations, we systematically investigated the catalytic behavior of Pt clusters on TiO2 under operational reaction conditions. We studied three distinct catalytic sitesPt cluster surfaces, oxygen vacancies (OV) on TiO2, and Pt–OV–Ti interfacesand revealed that the Pt–OV–Ti interface exhibited the highest RWGS activity via a redox mechanism. This synergy enhances CO2 activation and facilitates oxygen reduction more effectively than the isolated OV on TiO2, which show 4-fold lower activity. In contrast, CO-covered Pt clusters show minimal CO2 activation but serve as H2 dissociation sites, enabling hydrogen spillover to adjacent OV on TiO2, thereby sustaining the RWGS process. Kinetic analysis revealed OH reduction to H2O as the rate-determining step on both interfacial Pt–OV–Ti and at the OV on the TiO2–X support. These findings highlight the pivotal role of the Pt–OV–Ti interface in driving the RWGS and offer a design strategy for optimizing high-temperature CO2 hydrogenation catalysts by maximizing the number of interfacial active sites.
Keywords: Reverse water gas shift reaction, CO2 reduction, Pt/TiO2 , Theoretical modeling, Interfacial active sites, Redox mechanism, Strong metal support interaction


Introduction
The urgent need to mitigate climate change and reduce reliance on fossil fuels has driven intense research into CO2 conversion technologies. , The reverse water–gas shift (RWGS) reaction, which converts CO2 into CO via hydrogenation, is a critical step in carbon recycling, enabling the production of fuels and chemicals − through processes such as Fischer–Tropsch synthesis and the CAMERE (CO2 hydrogenation to methanol via the RWGS) process. However, the RWGS reaction is endothermic and requires high temperatures , to overcome thermodynamic limitations, making the development of efficient catalysts essential.
Metal catalysts supported on reducible oxides have been widely explored for RWGS because of their ability to activate CO2 through oxygen vacancies (OV) and metal–oxide interfacial interactions. − Among them, Pt/TiO2 catalysts demonstrate outstanding RWGS activity and selectivity, particularly in the 300–500 °C range, − owing to strong metal–support interactions (SMSIs). Highly dispersed Pt clusters (1–2 nm) are particularly promising, offering high atomic utilization, tunable electronic structures, and diverse coordination environments that enhance catalytic performance. While three potential active sitesPt cluster surfaces, OV on the TiO2 support, and Pt–OV–TiO2 interfacial siteshave been proposed, the precise nature of the active site and its role in the RWGS process remain debated. ,−
Experimental techniques such as XPS, IR, SEM, and XAS provide valuable insights into catalyst structure but face limitations in capturing dynamic surface phenomena and reaction intermediates under operando conditions. − Prior studies have proposed multiple RWGS mechanisms on Pt/TiO2, including a redox pathway involving both Pt sites and OV, − a carbonate-mediated mechanism at support vacancies, and a formate/bicarbonate pathway at Pt–OV–Ti3+ interfaces. However, the lack of consensus on these mechanistic pathways has hindered the establishment of the structure–activity relationship necessary for rational catalyst design.
In this study, we employ a multiscale theoretical approach to resolve the active sites and mechanism of RWGS on Pt/TiO2–X clusters. Ab initio thermodynamic analysis determines the stability of hydroxyl groups and OV on the TiO2–X support, whereas grand canonical Monte Carlo (GCMC) simulations identify the most stable Pt cluster structures under reaction conditions, with Pd8(CO)6 emerging as the most stable configuration. Microkinetic simulations, which are based on first-principles calculations, reveal that RWGS occurs predominantly via a redox mechanism at both the TiO2–OV and Pt–OV–Ti interfaces. Notably, the Pt–OV–Ti interfacial sites exhibit the highest intrinsic activity compared with the Pd8(CO)6 cluster surface and OV sites on TiO2–X , enhancing CO2 activation by lowering the dissociation barrier. This superior performance is attributed to its enhanced CO2 activation and efficient OV regeneration via OH reduction. The Pt cluster has minimal RWGS activity but serves as a H2 dissociation site, facilitating hydrogen spillover to OV sites on TiO2–X , where the reaction proceeds with 4-fold lower activity than the Pt–OV–Ti interface. Kinetic analysis further identified OH + H → H2O as the rate-determining step. These findings provide fundamental insights into RWGS catalysis and offer a design strategy for optimizing high-temperature CO2 hydrogenation catalysts by maximizing the number of interfacial active sites.
Results and Discussion
Oxygen Vacancy-Mediated RWGS on a TiO2–X Support
OV on TiO2 surfaces are well-known active sites for CO2 activation, facilitating the formation of key intermediates such as carbonates, bicarbonates, and formates. ,,, To assess their role in the RWGS reaction, we systematically investigated the surface structure and vacancy formation dynamics of TiO2 under reaction conditions. Using rutile TiO2(110), the most stable and frequently exposed surface, , we simulated its behavior in RWGS-relevant environments. At temperatures above 912 K and a partial H2 pressure of 0.67 bar, a 1/6 monolayer (ML) of surface oxygen atoms undergoes hydroxylation. As the temperature decreases, the entropy-driven effect of H2 diminishes, allowing a higher hydroxyl coverage of up to 1/3 ML (Figure a and Figure S1).
1.
Surface structures and PES for RWGS on TiO2(110). (a) Phase diagram of the OH coverage on the TiO2 (110) surface with a partial H2 pressure of 0.67 bar. (b) Gibbs formation energy of oxygen vacancies on TiO2 (110) with a hydroxylation coverage of 1/3 ML at 803 K and partial pressures of H2 and H2O of 0.67 bar and 0.1 mbar, respectively. (c) Configurations (above) and potential energy surface (below) for oxygen vacancy formation on the hydroxylated TiO2 (110) surface. (d) Reaction mechanism (above) and potential energy surface (below) for CO2 dissociation activation at the OV site on the hydroxylated TiO2–X (110) surface.
At 803 K and with H2 and H2O partial pressures of 0.67 bar and 0.10 mbar, respectively, a single OV per periodic p(2 × 4) unit cell forms on TiO2(110) with a formation energy of −0.01 eV. However, additional vacancy formation is thermodynamically endothermic, requiring a minimum energy input of 0.07 eV (Figure b and Figure S2). These vacancies originate from the recombination of adjacent hydroxyl groups, generating H2O and an oxygen defect with an activation barrier of 1.53 eV and an enthalpy of 1.44 eV. The desorption of H2O, with an energy requirement of 0.95 eV, stabilizes a significant low OV concentration under high-temperature RWGS conditions (Figure c).
Our calculations show that CO2 strongly chemisorbs at the OV on the TiO2 support, forming CO3 ## intermediates with an adsorption energy of −1.13 eV (Figure S3). # indicates the oxygen vacancy site on the TiO2 support. When hydroxylation occurs at the oxygen atoms adjacent to the OV, forming an OH# site, CO2 either weakly adsorbs or physically binds to the surface three-coordinated lattice oxygen, forming CO2 # and CO3’## intermediates with weaker adsorption energies of −0.29 eV and −0.22 eV, respectively. In contrast, in the absence of oxygen vacancies, the CO2 adsorption energy is much weaker at −0.14 eV. These findings emphasize the critical role of oxygen vacancies in the activation and adsorption of CO2, which is key to enhance the catalytic activity of the RWGS.
Furthermore, we systematically evaluated three CO2 hydrogenation pathways involving CO3 ##, CO3H##, and HCO2 # (Figure d). Our results indicate that CO formation is most favorable via the direct dissociation of CO3 ##, which has the lowest activation barrier of 1.70 eV (Figure d, Figure S4 and Table S1). The hydrogenation of CO3 ## to CO3H## followed by dissociation presents a slightly higher barrier of 1.75 eV, whereas the conversion of CO3H## to HCO2 # and its subsequent dissociation is the least favorable, requiring 2.81 eV. Therefore, these findings highlight that oxygen vacancies on TiO2 primarily facilitate RWGS via the CO3 ## and CO3H## intermediates, with the CO3 ## pathway being the most energetically efficient route for CO production.
Structures of Pt8/TiO2–X under RWGS Conditions
To comprehensively explore the nature of the active sites in Pt/TiO2 cluster-catalyzed RWGS in situ, we employed genetic algorithm (GA)-driven GCMC simulations to determine the Pt8/TiO2–X configuration under reaction conditions. We studied the configurations of Pt8 clusters (∼1 nm) on TiO2–X with an oxygen vacancy concentration of 1/24 ML. The ‘magic-numbered’ Pt8 cluster is chosen as a representative model owing to its exceptional stability , and size, which closely aligns with those of experimentally synthesized cluster catalysts. ,,,, In the most stable Pt8/TiO2–X structure, the OV is fully occupied by the Pt8 cluster and positioned centrally, as depicted in Figure a.
2.
In situ configuration of Pt 8 /TiO2–X under RWGS conditions. (a) Adsorption energy of CO and dissociative adsorption energy of H2 on the Pt8/TiO2–X cluster and the structures of Pt8/TiO2–X and Pt8(CO)6/TiO2–X identified by GA-driven GCMC simulations. (b) Adsorption free energy for different CO coverages on the Pt8/TiO2–X cluster at 803 K and a CO partial pressure of 0.13 bar. (c) Phase diagram of hydroxylation of the interfacial lattice oxygen of Pt8(CO)6/TiO2–X with a partial H2 pressure of 0.67 bar. (d) Adsorption free energy for varying coverage of H on the Pt8(CO)6/TiO2–X cluster (left) and free energy of oxygen vacancy formation at the Pt8(CO)6/TiO2–X interface (right) at 803 K and partial pressures of H2 and H2O of 0.67 bar and 0.1 mbar, respectively. The blue circle indicates the position of interfacial OV.
Among the species present in the reaction environment, H2 and CO adsorb strongly onto the Pt cluster surfaces, forming the predominant surface coverage species (Figure S5). Our calculations show that the dissociative adsorption energy of H2 on Pt8/TiO2–X clusters ranges from −0.96 to −0.81 eV, primarily occurring at Pt–Pt bridge sites. In contrast, CO adsorption on Pt is significantly stronger by 1.00–1.20 eV, predominantly occupying the top sites of the Pt atoms (Figure a and Figure S6). These findings indicate that, under reaction conditions, the Pt cluster surfaces are preferentially covered by CO, which could have implications for the reaction pathway and active site identification in the RWGS reaction.
GCMC simulations were conducted at 803 K and a CO partial pressure of 0.13 bar (resulting in 40% CO2 conversion at a total pressure of 1 bar and a H2:CO2 feed ratio of 2:1). As shown in Figure b, the Pt8/TiO2–X cluster stabilized with six CO molecules exhibited the lowest formation energy of −2.49 eV. The stability of the Pt8(CO)n clusters decreases with increasing CO coverage, suggesting coverage of approximately 3/4 ML on the Pt8 clusters under reaction conditions (Figures a, b, and Figure S7). In the stable Pt8(CO)6/TiO2–X structure, five CO molecules occupy the top sites of the Pt atoms, one CO molecule occupies a bridge site, and the Pt8 cluster relocates to its edge to cover the OV on the support.
We examined the hydroxyl species at the Pt8(CO)6/TiO2–X interface under reaction conditions (Figure c). Across the temperature range of 600–950 K, nearly all the two-coordinated oxygen atoms around the interface were hydroxylated, except for one lattice oxygen atom covered by the Pt clusters. This resulted in a slight decrease in surface hydroxyl coverage compared with that of the clean TiO2(110) surface, which decreased to 6/23 ML (Figure S8). Additionally, we explored hydrogen species adsorption on the Pt8 cluster surfaces and the distribution of OV at the interface. Our analysis revealed that although free sites exist on the Pt8(CO)6 cluster surfaces, hydrogen adsorption was thermodynamically unfavorable, with the lowest free energy for adsorbing three hydrogen atoms reaching 0.29 eV (Figure d and Figure S9). This aligns with experimental observations that hydrogen adsorption on Pt surfaces covered by carbonyl groups is challenging. Finally, our calculations demonstrate that the Pt8(CO)6/TiO2–X interface promotes oxygen vacancy formation by −0.14 eV, increasing the OV concentration to 1/8 ML by generating two adjacent bare oxygen vacancies (Figures d, S10–S11).
Mechanistic Insights into the RWGS over the Pt8(CO)6/TiO2–X System
We calculated the potential energy surface (PES) of RWGS at the interface of the most stable Pt8(CO)6/TiO2–X surface identified by the GCMC simulations under reaction conditions. First, we artificially desorbed a CO molecule from the surface of Pt8(CO)6 to form Pt8(CO)5, creating vacancies for the subsequent adsorption and dissociation of CO2 and H2. CO2 adsorption on the Pt cluster surface was weak, with an energy of only −0.31 eV (Figure S12). In contrast, at the Pt–OV–Ti interface, CO2 adsorption was significantly enhanced, reaching −1.81 eV (Figure S13 and Table S2), far surpassing the adsorption strength at oxygen vacancies on the TiO2–X support (Figure S3 and Table S1). This highlights the superior CO2 activation ability of the Pt–OV–Ti interfacial sites.
Once adsorbed at the Pt–OV–Ti interface, CO2 & undergoes either direct dissociation or hydrogenation to form intermediates such as COOH& or HCOO& (Figure a). & and * are the sites at Pt–OV–Ti and on the Pt cluster. Direct CO2 & dissociation is the most favorable process, with the lowest activation barrier (E a) of 0.43 eV (Figure b and Table S2). The oxygen fragment from CO2 & binds to the OV site, whereas the CO fragment is stabilized on Pt, facilitating CO2 & decomposition (Figure S13). Alternatively, the second pathway involves the hydrogenation of CO2 & to form the HCOO& intermediate, which then undergoes C–O bond cleavage to form HCO* under the synergistic action of Pt–OV–Ti interface, eventually dissociating to produce CO, with an apparent activation barrier of 1.06 eV (Figure c). This pathway is hindered primarily by the initial hydrogenation step, whereas the subsequent dissociation steps of HCOO& and HCO* are relatively facile, with lower activation barriers of 0.60 and 0.33 eV, respectively (Figure b, Figure S13, and Table S2). In contrast, the COOH& pathway, which involves the formation and dissociation of COOH& intermediates, presents the highest activation barrier of 1.78 eV, making it the least favorable (Figure c).
3.
RWGS mechanism study over the Pt8(CO)6/TiO2–X system. (a) Scheme of the mechanism of the RWGS reaction involving three distinct CO2 activations at the interface. (b) Potential energy surface for the RWGS at the Pt–OV–Ti interfaces. The black, blue and pink lines represent direct CO2 dissociation, the COOH pathway and the HCOO pathway, respectively. The elementary activation barriers are indicated in eV. (c) Overall activation barriers for CO2 activation at the Pt–OV–Ti interfaces. The black, blue and pink columns represent direct CO2 dissociation, the COOH pathway and the HCOO pathway, respectively. All the energies indicated are in eV. (d) TOF for RWGS and coverage for the intermediates CO2, H and OH at the Pt–OV–Ti interface and hydroxylated TiO2–X (110) surface at 803 K, 1 bar, and a feed ratio of H2:CO2 of 2:1. (e) Potential energy surface for H2 adsorption and migration on a Pt cluster (blue line) and spillover to the TiO2–X support (green line). All the energies indicated are in eV.
Our microkinetic simulations, which are based on the calculated PES, provide a detailed assessment of the RWGS activity at the Pt–OV–Ti interface and TiO2–X support. At 803 K, a total pressure of 1 bar, and a H2:CO2 ratio of 2:1, the turnover frequency (TOF) at the Pt–OV–Ti interface is 2351 s–1, which is four times greater than the value of 570 s–1 observed for the TiO2–X support (Figure d). Consistent with our PES analysis, the microkinetic simulations revealed that the optimal CO2 activation pathway at these two sites involves direct CO2 dissociation and CO3 ## dissociation, respectively. The degree of rate control (DRC) analysis indicates that the rate-determining step (RDS) at both active sites is the reduction of OH# to generate H2O#. The activation barriers for this elementary step are 0.35 eV at the Pt–OV–Ti interface and 1.53 eV at the OV sites on the TiO2–X support (Figures b and d), respectively. Consequently, the Pt–OV–Ti interface dominated the RWGS activity, whereas TiO2–X support has a less contribution to the overall activity.
The coverage of key intermediates, such as CO2 &, CO3 ##, H*, H# and OH#, provides crucial insight into the differing reactivities between the Pt–OV–Ti interface and the TiO2–X support. Specifically, the coverage of CO2 & at the Pt–OV–Ti interface is 10–7, which is significantly greater than the CO3 ## coverage of 10–9 on the TiO2–X support, confirming that the interfacial Pt–OV–Ti sites greatly enhance CO2 adsorption and activation. However, the lower equilibrium coverage of H* on the Pt clusters and OH# at the interface, compared with the higher coverage of H# and OH# on TiO2–X (where H# refers to H adsorbed on the two-coordinated lattice oxygen adjacent to the OV, Figure d), mitigates the activity advantage of the Pt–OV–Ti interface, narrowing the TOF difference between the two structures.
While metallic Pt readily facilitates H2 dissociation, our calculations confirm that this process is barrierless and strongly exothermic (−1.05 eV) (Figure S14). However, owing to strong competitive adsorption by CO and the optimal site for subsequent CO2 adsorption at the interface, H* must migrate to the remaining free Pt sites. This migration increases the H2 dissociative adsorption energy on the Pt8 cluster to −0.29 eV (Figure e and Figure S15), leading to low H* coverage on the Pt surface. Consequently, the reduction of O# or OH# at the interface is hindered. As a result, although OH# hydrogenation to H2O# has a lower activation barrier (0.35 eV) than CO2 dissociation (0.43 eV) and O# reduction (0.87 eV), the OH# reduction step becomes kinetically limiting due to its low coverage, thus governing the overall rate.
On the TiO2–X support, hydrogen for the RWGS reaction originates from both direct H2 dissociation and spillover from Pt clusters. The apparent activation barriers for these processes are 1.57 and 1.46 eV, respectively, with a net energy release of −1.30 eV (Figure e and Figures S16 and S17). This strong exothermicity led to high H# coverage on TiO2–X , facilitating OH# reduction. Consequently, despite weaker CO2 adsorption, a greater CO2 dissociation barrier, and a greater activation barrier for the RDS on TiO2–X , its intrinsic RWGS activity is just 4-fold lower than that at the Pt–OV–Ti interface.
Our theoretical simulations are consistent with experimental studies on Pt/r-TiO2 clusters in RWGS, combining DFT, KMC simulations, experimental testing, and in situ FTIR characterization, , which confirm that the Pt–OV–Ti3+ interface is the active site. The reported TOF of 6.3 s–1 for Pt/r-TiO2 (Pt size ∼ 1 nm) at 623 K and a H2:CO2 ratio of 2:1 closely matches our simulation result of 10.4 s–1 under the same conditions. Further Arrhenius analysis revealed an activation energy of 0.80 eV for CO formation on 0.1%Pt/r-TiO2 at 250–400 °C, which deviates slightly from our calculated 1.11 eV barrier from microkinetic simulations. These experimental results validate our theoretical model. While the rate-determining step for Pt/TiO2-catalyzed RWGS remains debated, our integrated structural and microkinetic analysis identifies the OH+H→H2O step as the rate-determining step, offering new insights into RWGS catalysis.
Origins of Superior Catalytic Performance over Pt–OV–Ti Interfacial Sites
In first-principles microkinetic simulations, we evaluated three distinct catalytic sites for the RWGS reaction under reaction conditions: the TiO2–X support, the Pt/TiO2–X interface, and the Pt cluster surfaces. Our results highlight that the Pt–OV–Ti interfacial site has the highest intrinsic catalytic activity (Figure d). This enhanced performance stems from the significant promotion of CO2 activation and the efficient reduction of OH# species at the interface.
The catalyst surface facilitates electron transfer to CO2, inducing a rearrangement of its molecular orbitals and disrupting the linear CO2 configuration. This distortion weakens the CO bond, thereby enhancing adsorption activation. ,, On the TiO2–X support and Pt clusters, CO2 adsorption is weaker, with adsorption energies of −1.13 eV and −0.31 eV, respectively. Bader charge analysis revealed minimal charge transfer to CO2, with values of −0.24 |e| and −0.39 |e|, respectively (Figure a). Specifically, CO2 adsorption on Pt clusters leads to slight bending of the molecule, resulting in an O–C–O bond angle of 143.7°, which further decreases to 125.6° at the OV on the TiO2–X support. In contrast, at the Pt–OV–Ti interface, charge transfer to CO2 is significantly greater (−0.79 |e|), with a corresponding adsorption energy of −1.81 eV, and the O–C–O bond angle decreases to 121.4°, indicating more activated CO2 molecules.
4.
Electronic structure analysis of adsorbed intermediates and active sites. (a) Structures of CO2 adsorbed on the TiO2–X (110) support, Pt–OV–Ti interface and Pt cluster surface sites and their corresponding isosurfaces of differential charge density with isovalues of 0.02 |e|/Å3. Yellow and cyan represent charge accumulation and depletion, respectively. (b) Density of state (DOS) plot of CO2 molecules and bonded Ti, O and Pt atoms at oxygen vacancies on the TiO2–X (110) support, interfacial Pt–OV–Ti and Pt cluster surfaces, respectively. (c) Projected crystal orbital Hamilton population (pCOHP) plot of O–Ti bonds for the initial and transition states of OH reduced to H2O by H at the oxygen vacancy on the TiO2–X (110) support and interfacial Pt–OV–Ti sites, respectively. (d) Electron density map of the TiO2–X (110) surface and Pt–OV–Ti interface.
From a molecular orbital perspective, on the TiO2–X support, the s and p orbitals of CO2 strongly hybridize with the d orbitals of Ti atoms at oxygen vacancies below the Fermi level at −9.00, −7.50, and −3.60 eV (Figure b). Additionally, CO2 forms strong chemical bonds with the lattice oxygen atoms adjacent to OV. This hybridization, rather than tiny charge transfer, explains the relatively strong adsorption energy of −1.13 eV. In contrast, the interaction of CO2 with Pt clusters is weak, with limited molecular orbital hybridization below −6.00 eV, which underscores the poor adsorption and activation of CO2 by Pt clusters, rendering them less effective as active sites for the RWGS reaction. At the Pt–OV–Ti interface, however, CO2 undergoes both hybridization with the d orbitals of Ti atoms at the OV site and strong overlapping interactions with the d orbitals of metallic Pt atoms. This synergy at the Pt–OV interface significantly enhances CO2 adsorption and activation, facilitating more efficient CO2 dissociation.
Moreover, the reduction of OH# species left behind by CO2 dissociationidentified as the RDS by microkinetic analysisexhibits significant differences between the OV sites on the TiO2–X support and the interfacial Pt–OV–Ti active sites. Projected crystal orbital Hamilton population (pCOHP) analysis demonstrated that the Ti–O bond formed by the OH# fragment at OV sites on the TiO2–X support was weakened in the transition state compared with the initial state, with a -pCOHP integral of 1.88. In contrast, at the Pt–OV–Ti interface, the O–Ti chemical bond is significantly stronger, with an -IpCOHP of 2.60 (Figure c). Furthermore, the extent of bond weakening relative to the initial Ti–O bonding strength is significantly enhanced at the Pt–OV–Ti sites, resulting in a lower reaction barrier at the interface. In summary, the interfacial synergistic sites of Pt–OV–Ti exhibit stronger binding abilities for both CO2 and OH#, facilitating the RWGS reaction. This enhancement is due to the SMSI between the Pt clusters and the TiO2–X support, which promotes the generation of oxygen vacancies and results in greater charge accumulation at the interface than does the bare TiO2–X support (Figure d). Consequently, the adsorption of key species and intermediates is stronger, the reaction PES is lower, and the intrinsic catalytic activity is increased.
Conclusions
In this study, we investigated the active sites of Pt/TiO2 cluster catalysts for the RWGS reaction, focusing on the TiO2–X support, Pt/TiO2–X interface, and Pt clusters. GA-driven GCMC simulations revealed the stable Pt8(CO)6/TiO2–X configuration under reaction conditions, followed by first-principles calculations and microkinetic analysis. Our findings reveal that Pt–OV–Ti interfacial sites exhibit the highest intrinsic activity (TOF = 2351 s–1), which is attributed to efficient CO2 activation and oxygen reduction. Although oxygen vacancies on TiO2–X also contribute (TOF = 570 s–1), their activity is 4-fold lower. In contrast, Pt clusters show negligible CO2 activation capacity, but serve as H2 dissociation sites, facilitating hydrogen spillover to TiO2–X for RWGS. The strong metal–support interaction (SMSI) between the Pt clusters and TiO2–X promotes the formation of Pt–OV–Ti sites, further enhancing CO2 activation and oxygen reduction. These insights underscore the critical role of Pt–OV–Ti interfaces in driving efficient CO2 conversion and provide a strategy for optimizing metal–oxide interfaces to enhance RWGS performance.
Computational Methods
Spin-polarized density functional theory (DFT) calculations were conducted via the Vienna Ab initio Simulation Package (VASP) , with the projector augmented wave (PAW) potential. The generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof exchange–correlation functional was adopted to describe the electron exchange–correlation interaction. A kinetic energy cutoff of 400 eV was applied. Convergence criteria for electronic optimization were set at 10–4 eV/atom, and geometry optimizations were deemed converged when the forces on each ion were less than 0.05 eV/Å. A p(4 × 2) supercell slab with a two-layer oxide thickness was used to simulate the (110) surface of rutile TiO2. The supported Pt8 clusters were identified via a GA, and their configurations under reaction conditions were determined via GCMC simulations. A vacuum of 15 Å was used to separate neighboring slabs to avoid dipole effects between them. A Monkhorst–Pack k-point mesh of size 2 × 2 × 1 was used to sample the Brillouin zones. GGA+U correction was used when the strongly correlated d electrons of Ti with U eff were 4.2.
During structure optimization, the adsorbates, the metal cluster and the topmost oxide are relaxed. The adsorption energies (E ads) were calculated via the following equation:
| 1 |
where E total, E slab, and E molecule(g) are the total energies of the adsorbed system, clean slab and adsorbate in the gas phase, respectively. The activation barrier (E a) and reaction energy (E r) were calculated as
| 2 |
where E IS, E FS and E TS are the total energy of the initial state, final state, and corresponding transition state, respectively. The transition states (TSs) were identified via the improved force reversal method until all the forces were less than 0.05 eV/Å. Some of the TSs were confirmed by climbing-image nudged elastic band (CI-NEB) and dimer methods.
The turnover frequency (TOF), DRC, coverage of intermediates, and reaction mechanism were analyzed via CATKINAS software. The simulations were conducted at a reaction temperature of 803 K, with an initial reactant feed ratio of CO2 to H2 of 1:2 and a total pressure of 1 bar. These conditions are consistent with those reported in the literature. The detailed methodologies used for the microkinetic analysis, GA, and GCMC can be found in the Supporting Information.
Supplementary Material
Acknowledgments
This work was supported by the Key Technologies R&D Program of China (2021YFA1502804), the Strategic Priority Research Program of the Chinese Academy of Science (XDB0450102), the National Natural Science Foundation of China (22172150, 22222306, 22221003, and 22432004), the Innovation Program for Quantum Science and Technology (2021ZD0303302), and the high-performance computational resources provided by the University of Science and Technology of China and Hefei Advanced Computing Center.
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/prechem.5c00010.
Detailed methodology, DFT calculation structures, potential energy surfaces and calculated energetics (PDF)
Jin-Xun Liu led the conceptualization and design of the DFT calculations. Feng Li contributed to the DFT calculations and data analysis. All the authors participated in writing the manuscript and in the overall scientific interpretation.
The authors declare no competing financial interest.
References
- Doney S. C., Fabry V. J., Feely R. A., Kleypas J. A.. Ocean Acidification: The Other CO2 Problem. Annu. Rev. Mar. Sci. 2009;1(1):169–192. doi: 10.1146/annurev.marine.010908.163834. [DOI] [PubMed] [Google Scholar]
- Rahman F. A., Aziz M. M. A., Saidur R., Bakar W. A. W. A., Hainin M. R., Putrajaya R., Hassan N. A.. Pollution to Solution: Capture and Sequestration of Carbon Dioxide (CO2) and Its Utilization as a Renewable Energy Source for a Sustainable Future. Renew. Sustain. Energy Rev. 2017;71:112–126. doi: 10.1016/j.rser.2017.01.011. [DOI] [Google Scholar]
- Nielsen D. U., Hu X.-M., Daasbjerg K., Skrydstrup T.. Chemically and Electrochemically Catalyzed Conversion of CO2 to CO with Follow-Up Utilization to Value-Added Chemicals. Nat. Catal. 2018;1(4):244–254. doi: 10.1038/s41929-018-0051-3. [DOI] [Google Scholar]
- Ahmadi Khoshooei M., Wang X., Vitale G., Formalik F., Kirlikovali K. O., Snurr R. Q., Pereira-Almao P., Farha O. K.. An Active, Stable Cubic Molybdenum Carbide Catalyst for the High-Temperature Reverse Water-Gas Shift Reaction. Science. 2024;384(6695):540–546. doi: 10.1126/science.adl1260. [DOI] [PubMed] [Google Scholar]
- Ye R.-P., Ding J., Gong W., Argyle M. D., Zhong Q., Wang Y., Russell C. K., Xu Z., Russell A. G., Li Q.. et al. CO2 Hydrogenation to High-Value Products via Heterogeneous Catalysis. Nat. Commun. 2019;10(1):5698. doi: 10.1038/s41467-019-13638-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang X., Nie X., Guo X., Song C., Chen J. G.. Recent Advances in Carbon Dioxide Hydrogenation to Methanol via Heterogeneous Catalysis. Chem. Rev. 2020;120(15):7984–8034. doi: 10.1021/acs.chemrev.9b00723. [DOI] [PubMed] [Google Scholar]
- Callaghan, C. A. Kinetics and Catalysis of the Water-Gas-Shift Reaction: A Microkinetic and Graph Theoretic Approach. Thesis; Worcester Polytechnic Institute, 2006. [Google Scholar]
- Wang W., Wang S., Ma X., Gong J.. Recent Advances in Catalytic Hydrogenation of Carbon Dioxide. Chem. Soc. Rev. 2011;40(7):3703–3727. doi: 10.1039/c1cs15008a. [DOI] [PubMed] [Google Scholar]
- Kattel S., Liu P., Chen J. G.. Tuning Selectivity of CO2 Hydrogenation Reactions at the Metal/Oxide Interface. J. Am. Chem. Soc. 2017;139(29):9739–9754. doi: 10.1021/jacs.7b05362. [DOI] [PubMed] [Google Scholar]
- Belgamwar R., Verma R., Das T., Chakraborty S., Sarawade P., Polshettiwar V.. Defects Tune the Strong Metal-Support Interactions in Copper Supported on Defected Titanium Dioxide Catalysts for CO2 Reduction. J. Am. Chem. Soc. 2023;145(15):8634–8646. doi: 10.1021/jacs.3c01336. [DOI] [PubMed] [Google Scholar]
- Kang H., Zhu L., Li S., Yu S., Niu Y., Zhang B., Chu W., Liu X., Perathoner S., Centi G.. Generation of Oxide Surface Patches Promoting H-Spillover in Ru/(TiOx)MnO Catalysts Enables CO2 Reduction to CO. Nat. Catal. 2023;6(11):1062–1072. doi: 10.1038/s41929-023-01040-0. [DOI] [Google Scholar]
- Liu H.-X., Li S.-Q., Wang W.-W., Yu W.-Z., Zhang W.-J., Ma C., Jia C.-J.. Partially Sintered Copper-Ceria as Excellent Catalyst for the High-Temperature Reverse Water Gas Shift Reaction. Nat. Commun. 2022;13(1):867. doi: 10.1038/s41467-022-28476-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang C.-X., Liu H.-X., Gu H., Li J.-Y., Lai X.-M., Fu X.-P., Wang W.-W., Fu Q., Wang F. R., Ma C.. Hydroxylated TiO2-Induced High-Density Ni Clusters for Breaking the Activity-Selectivity Trade-Off of CO2 Hydrogenation. Nat. Commun. 2024;15(1):8290. doi: 10.1038/s41467-024-52547-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nelson N. C., Nguyen M.-T., Glezakou V.-A., Rousseau R., Szanyi J.. Carboxyl Intermediate Formation via an In Situ-Generated Metastable Active Site During Water-Gas Shift Catalysis. Nat. Catal. 2019;2(10):916–924. doi: 10.1038/s41929-019-0343-2. [DOI] [Google Scholar]
- Nelson N. C., Szanyi J.. Heterolytic Hydrogen Activation: Understanding Support Effects in Water-Gas Shift, Hydrodeoxygenation, and CO Oxidation Catalysis. ACS Catal. 2020;10(10):5663–5671. doi: 10.1021/acscatal.0c01059. [DOI] [Google Scholar]
- Torres-Sempere G., Blay-Roger R., Luque-Álvarez L. A., Santos J. L., Bobadilla L. F., Pastor-Pérez L., Centeno M. A., Hernández W. Y., Yousef I., Odriozola J. A.. Subnanometric Pt Clusters Dispersed over Cs-Doped TiO2 for CO2 Upgrading via Low-Temperature RWGS: Operando Mechanistic Insights to Guide an Optimal Catalyst Design. J. Mater. Chem. A. 2024;12(3):1779–1792. doi: 10.1039/D3TA05482A. [DOI] [Google Scholar]
- Kim S. S., Park K. H., Hong S. C.. A Study of the Selectivity of the Reverse Water-Gas-Shift Reaction over Pt/TiO2 Catalysts. Fuel Process. Technol. 2013;108:47–54. doi: 10.1016/j.fuproc.2012.04.003. [DOI] [Google Scholar]
- 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. doi: 10.1016/j.apcata.2012.02.021. [DOI] [Google Scholar]
- 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. doi: 10.1016/j.apcatb.2012.02.023. [DOI] [Google Scholar]
- Kattel S., Yan B., Chen J. G., Liu P.. CO2 Hydrogenation on Pt, Pt/SiO2 and Pt/TiO2: Importance of Synergy Between Pt and Oxide Support. J. Catal. 2016;343:115–126. doi: 10.1016/j.jcat.2015.12.019. [DOI] [Google Scholar]
- Chen X., Su X., Duan H., Liang B., Huang Y., Zhang T.. 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. doi: 10.1016/j.cattod.2016.03.020. [DOI] [Google Scholar]
- Zhang H., Xie L., Huang C., Ren Z., Wang H., Hu J., Zhang H., Jiang Z., Song F.. Exploring the CO2 Reduction Reaction Mechanism on Pt/TiO2 with the Ambient-Pressure X-ray Photoelectron Spectroscopy. Appl. Surf. Sci. 2021;568:150933. doi: 10.1016/j.apsusc.2021.150933. [DOI] [Google Scholar]
- Chen Z., Liang L., Yuan H., Liu H., Wu P., Fu M., Wu J., Chen P., Qiu Y., Ye D.. Reciprocal Regulation Between Support Defects and Strong Metal-Support Interactions for Highly Efficient Reverse Water Gas Shift Reaction Over Pt/TiO2 Nanosheets Catalysts. Appl. Catal. B: Environ. 2021;298:120507. doi: 10.1016/j.apcatb.2021.120507. [DOI] [Google Scholar]
- Lin L.-Y., Kavadiya S., He X., Wang W.-N., Karakocak B. B., Lin Y.-C., Berezin M. Y., Biswas P.. Engineering Stable Pt Nanoparticles and Oxygen Vacancies on Defective TiO2 via Introducing Strong Electronic Metal-Support Interaction for Efficient CO2 Photoreduction. Chem. Eng. J. 2020;389:123450. doi: 10.1016/j.cej.2019.123450. [DOI] [Google Scholar]
- Liu H.-X., Li J.-Y., Qin X., Ma C., Wang W.-W., Xu K., Yan H., Xiao D., Jia C.-J., Fu Q.. Ptn-Ov Synergistic Sites on MoOx/γ-Mo2N Heterostructure for Low-Temperature Reverse Water-Gas Shift Reaction. Nat. Commun. 2022;13(1):5800. doi: 10.1038/s41467-022-33308-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang R., Wang X., Wang K., Wang H., Liu L., Wu X., Geng B., Chu X., Song S., Zhang H.. Synergism of Ultrasmall Pt Clusters and Basic La2O2CO3 Supports Boosts the Reverse Water Gas Reaction Efficiency. Adv. Energy Mater. 2023;13(12):2203806. doi: 10.1002/aenm.202203806. [DOI] [Google Scholar]
- Li W., Gan J., Liu Y., Zou Y., Zhang S., Qu Y.. Platinum and Frustrated Lewis Pairs on Ceria as Dual-Active Sites for Efficient Reverse Water-Gas Shift Reaction at Low Temperatures. Angew. Chem. 2023;135(37):e202305661. doi: 10.1002/ange.202305661. [DOI] [PubMed] [Google Scholar]
- Krishna D. N. G., Philip J.. Review on Surface-Characterization Applications of X-ray Photoelectron Spectroscopy (XPS): Recent Developments and Challenges. Appl. Surf. Sci. Adv. 2022;12:100332. doi: 10.1016/j.apsadv.2022.100332. [DOI] [Google Scholar]
- Hess C.. New Advances in Using Raman Spectroscopy for the Characterization of Catalysts and Catalytic Reactions. Chem. Soc. Rev. 2021;50(5):3519–3564. doi: 10.1039/D0CS01059F. [DOI] [PubMed] [Google Scholar]
- Tao F., Crozier P. A.. Atomic-Scale Observations of Catalyst Structures Under Reaction Conditions and During Catalysis. Chem. Rev. 2016;116(6):3487–3539. doi: 10.1021/cr5002657. [DOI] [PubMed] [Google Scholar]
- Su D. S., Zhang B., Schlögl R.. Electron Microscopy of Solid Catalysts Transforming from a Challenge to a Toolbox. Chem. Rev. 2015;115(8):2818–2882. doi: 10.1021/cr500084c. [DOI] [PubMed] [Google Scholar]
- Bobadilla L. F., Santos J. L., Ivanova S., Odriozola J. A., Urakawa A.. Unraveling the Role of Oxygen Vacancies in the Mechanism of the Reverse Water-Gas Shift Reaction by Operando DRIFTS and Ultraviolet-Visible Spectroscopy. ACS Catal. 2018;8(8):7455–7467. doi: 10.1021/acscatal.8b02121. [DOI] [Google Scholar]
- Ramamoorthy M., Vanderbilt D., King-Smith R. D.. First-Principles Calculations of the Energetics of Stoichiometric TiO2 Surfaces. Phys. Rev. B. 1994;49(23):16721. doi: 10.1103/PhysRevB.49.16721. [DOI] [PubMed] [Google Scholar]
- Pang C. L., Lindsay R., Thornton G.. Structure of Clean and Adsorbate-Covered Single-Crystal Rutile TiO2 Surfaces. Chem. Rev. 2013;113(6):3887–3948. doi: 10.1021/cr300409r. [DOI] [PubMed] [Google Scholar]
- Knight W., Clemenger K., De Heer W. A., Saunders W. A., Chou M., Cohen M. L.. Electronic Shell Structure and Abundances of Sodium Clusters. Phys. Rev. Lett. 1984;52(24):2141. doi: 10.1103/PhysRevLett.52.2141. [DOI] [Google Scholar]
- Guo B., Kerns K., Castleman A. Jr. Ti8C12+-Metallo-Carbohedrenes: a New Class of Molecular Clusters? Science. 1992;255(5050):1411–1413. doi: 10.1126/science.255.5050.1411. [DOI] [PubMed] [Google Scholar]
- Li S., Xu Y., Chen Y., Li W., Lin L., Li M., Deng Y., Wang X., Ge B., Yang C.. Tuning the Selectivity of Catalytic Carbon Dioxide Hydrogenation over Iridium/Cerium Oxide Catalysts with a Strong Metal-Support Interaction. Angew. Chem. 2017;129(36):10901–10905. doi: 10.1002/ange.201705002. [DOI] [PubMed] [Google Scholar]
- Have I. C. t., Kromwijk J. J., Monai M., Ferri D., Sterk E. B., Meirer F., Weckhuysen B. M.. Uncovering the Reaction Mechanism Behind CoO as Active Phase for CO2 Hydrogenation. Nat. Commun. 2022;13(1):324. doi: 10.1038/s41467-022-27981-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ko J., Kim B.-K., Han J. W.. Density Functional Theory Study for Catalytic Activation and Dissociation of CO2 on Bimetallic Alloy Surfaces. J. Phys. Chem. C. 2016;120(6):3438–3447. doi: 10.1021/acs.jpcc.6b00221. [DOI] [Google Scholar]
- Kresse G., Furthmüller J.. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B. 1996;54(16):11169–11186. doi: 10.1103/PhysRevB.54.11169. [DOI] [PubMed] [Google Scholar]
- Kresse G., Furthmüller J.. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996;6(1):15–50. doi: 10.1016/0927-0256(96)00008-0. [DOI] [Google Scholar]
- Blöchl P. E.. Projector Augmented-Wave Method. Phys. Rev. B. 1994;50(24):17953. doi: 10.1103/PhysRevB.50.17953. [DOI] [PubMed] [Google Scholar]
- Perdew J. P., Burke K., Ernzerhof M.. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996;77(18):3865. doi: 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- Morgan B. J., Watson G. W.. A DFT+ U Description of Oxygen Vacancies at the TiO2 Rutile (110) Surface. Surf. Sci. 2007;601(21):5034–5041. doi: 10.1016/j.susc.2007.08.025. [DOI] [Google Scholar]
- Sun K., Zhao Y., Su H.-Y., Li W.-X.. Force Reversed Method for Locating Transition States. Theor. Chem. Acc. 2012;131(2):1118. doi: 10.1007/s00214-012-1118-x. [DOI] [Google Scholar]
- Henkelman G., Uberuaga B. P., Jónsson H.. A Climbing Image Nudged Elastic Band Method for Finding Saddle Points and Minimum Energy Paths. J. Chem. Phys. 2000;113(22):9901–9904. doi: 10.1063/1.1329672. [DOI] [Google Scholar]
- Henkelman G., Jónsson H.. A Dimer Method for Finding Saddle Points on High Dimensional Potential Surfaces Using Only First Derivatives. J. Chem. Phys. 1999;111(15):7010–7022. doi: 10.1063/1.480097. [DOI] [Google Scholar]
- Chen J., Jia M., Hu P., Wang H.. CATKINAS: a Large-Scale Catalytic Microkinetic Analysis Software for Mechanism Auto-Analysis and Catalyst Screening. J. Comput. Chem. 2021;42(5):379–391. doi: 10.1002/jcc.26464. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




