Next Article in Journal
Faster Microwave-Assisted Synthesis of Microspherical Carbons from Commercial and Biomass-Derived Carbohydrates
Previous Article in Journal
From Pigment to Photocatalyst: CdSe/CdS Solutions Mimicking Cadmium Red for Visible-Light Dye Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Atomic-Scale Insights into CO2 and H2O Co-Adsorption on Sr2Fe1.5Mo0.5O6 Surfaces: Role of Electronic Structure and Dual-Site Interactions

1
Foshan Power Supply Bureau of Guangdong Power Grid, Foshan 528000, China
2
Institute of Science and Education Development, Xi’an Jiaotong University, 28 Xianning West Road, Xi’an 710049, China
3
School of Electronics and Information, Xi’an Polytechnic University, 19 Jinhua South Road, Xi’an 710048, China
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(9), 884; https://doi.org/10.3390/catal15090884
Submission received: 17 August 2025 / Revised: 8 September 2025 / Accepted: 12 September 2025 / Published: 15 September 2025
(This article belongs to the Special Issue Catalytic Conversion of CO2 or CO)

Abstract

Co-electrolysis of CO2 and H2O offers a promising route for efficient and controllable syngas production from greenhouse gases and water. However, the atomic-scale reaction mechanism remains elusive, especially on complex oxide surfaces. In this study, we employ density functional theory (DFT) to investigate the adsorption and activation of CO2 and H2O on the FeMoO-terminated (001) surface of Sr2Fe1.5Mo0.5O6 (SFM), a double perovskite of growing interest for solid oxide electrolysis. Our results show that CO2 strongly interacts with surface lattice oxygen, adopting a bent configuration with substantial charge transfer. In contrast, H2O binds more weakly at Mo sites through predominantly electrostatic interactions. Co-adsorption analyses reveal a bidirectional interplay: pre-adsorbed H2O enhances CO2 binding by altering its adsorption geometry, whereas pre-adsorbed CO2 weakens H2O adsorption due to competitive site occupation. This balance suggests that moderate co-adsorption may facilitate proton–electron coupling, while excessive coverage of either species suppresses activation of the other. Bader charge analysis, charge density differences, and projected density of states highlight the key role of Fe/Mo–O hybridized states near the Fermi level in mediating surface reactivity. These results, obtained for a perfect defect-free surface, provide a theoretical benchmark for disentangling intrinsic molecule–surface and molecule–molecule interactions, and offer guidance for designing high-performance perovskite electrocatalysts for CO2 + H2O co-electrolysis.

Graphical Abstract

1. Introduction

The electrochemical co-reduction of CO2 and H2O in solid oxide electrolysis cells (SOECs) represents a promising strategy for sustainable syngas (CO + H2) production, bridging the gap between renewable energy sources and synthetic fuels. Compared to conventional water electrolysis, the co-electrolysis approach offers enhanced carbon utilization, direct coupling to carbon capture, and greater thermodynamic efficiency at high operating temperatures, typically in the range of 700–900 °C [1,2,3]. These conditions not only reduce the kinetic barriers for gas activation but also allow reversible operation in solid oxide fuel cell (SOFC) and SOEC modes, providing system-level flexibility [4]. Moreover, syngas generated via co-electrolysis can be further processed through Fischer–Tropsch synthesis or methanol production, forming the basis for renewable liquid fuels and value-added chemicals [5,6,7,8]. In this process, the initial adsorption and activation of H2O and CO2 on the cathode surface govern the rates of the electrochemical reactions, so understanding these surface steps is critical for improving overall performance.
Traditionally, Ni-based electrodes have been the dominant choice in SOECs due to their high electronic conductivity and catalytic activity. However, Ni catalysts suffer from several issues including carbon deposition (coking), sulfur poisoning, and mechanical instability under redox cycling [9,10]. These limitations have motivated the exploration of alternative oxide-based electrodes, especially mixed ionic-electronic conductors (MIECs) with perovskite or double-perovskite structures [11]. In this context, the double perovskite Sr2Fe1.5Mo0.5O6 (SFM) has emerged as a particularly promising candidate. SFM exhibits favorable redox stability, high electrical conductivity, and fast oxygen-ion transport under reducing conditions, making it well-suited for the harsh environment of SOEC cathodes [12,13,14]. The synergistic interaction between Fe and Mo cations in the B-site lattice of SFM promotes oxygen vacancy formation and electron delocalization, which can enhance surface reactivity toward gas-phase molecules such as CO2 and H2O [15]. Furthermore, the Fe-centered electronic states provide redox-active sites, while the structural flexibility of the perovskite lattice enables compositional tuning through A-site or B-site doping to optimize adsorption and catalytic behavior [16].
Despite these advantages, the fundamental surface mechanisms by which CO2 and H2O adsorb and interact on SFM remain insufficiently understood. Experimental studies have provided indirect evidence of gas conversion performance and long-term stability, yet lack the atomic-scale resolution needed to unravel the initial activation steps. Here, we emphasize that our DFT calculations are performed on a defect-free, idealized SFM surface, which serves as a theoretical benchmark to systematically explore adsorption energetics and electronic interactions [17,18,19,20,21]. While real materials may contain vacancies or mixed-valence cations, the perfect-surface model provides a clear reference point for understanding fundamental molecule–surface interactions. Previous First-principles density functional theory (DFT) studies have shown that oxygen vacancies can strongly modify adsorption energetics and surface redox chemistry [22,23]. Therefore, our defect-free surface results should be viewed as a theoretical benchmark, with future work extending to defective surfaces to bridge with realistic SOEC conditions. Within this framework, a key open question is how the simultaneous presence of CO2 and H2O influences charge redistribution, adsorption energetics, and electronic structure on SFM surfaces. DFT simulations offer a powerful tool to fill this gap by quantifying adsorption energies, surface charge transfer, and electronic reorganization in a controlled and systematic manner [24,25,26,27].
In this study, we focus on the FeMoO-terminated (001) surface of Sr2Fe1.5Mo0.5O6 to explore the detailed adsorption configurations of CO2 and H2O, both individually and in co-adsorption states. We employ spin-polarized DFT calculations with Hubbard U corrections to account for the strong correlation effects of the transition metal d-orbitals, ensuring reliable treatment of the surface electronic structure. By computing adsorption energies, differential charge densities, Bader charges, and partial density of states (PDOS), we elucidate how the interplay between adsorbates and surface cations modulates the local chemical environment. Our results reveal competitive and cooperative interactions between CO2 and H2O during co-adsorption, providing insight into the initial stages of CO and H2 evolution. This study contributes to a molecular-level understanding of syngas generation on SFM and provides a theoretical foundation for future design of high-performance perovskite-based cathodes for CO2 + H2O co-electrolysis.

2. Results and Discussion

In this study, we constructed and optimized the geometry of the Sr2Fe1.5Mo0.5O6 (SFM) perovskite structure, as shown in Figure 1. In the optimized unit cell, Fe atoms are represented in brown, Mo in purple, O in red, and Sr in green. The calculated bond lengths of Fe-O and Mo-O in pristine SFM are 1.99 Å and 1.91 Å, respectively. A comparison of the lattice constants obtained in this work with those reported in the previous literature is provided in Table 1. The optimized parameters, with a = 5.542 Å and c = 7.940 Å, are in excellent agreement with literature values, confirming the reliability of our computational model for subsequent investigations.
Based on the optimized bulk structure, a 2 × 3 × 1 supercell was constructed, followed by cleaving along the (001) plane to model the surface. The surface slab contains 120 atoms and includes a 15 Å vacuum layer to simulate the molecular adsorption environment. The bottom two atomic layers were fixed to mimic the bulk environment, while the top two layers were fully relaxed. Structural relaxation revealed moderate adjustments in the Fe-O and Mo-O bond lengths at the surface. Figure 2 illustrates the atomic layer stacking along the (001) orientation of the optimized surface. The (001) direction of the SFM crystal contains alternating SrO, FeO2, and FeMoO layers (Figure S1). Among these, the difference in oxidation states between Fe and Mo (Fe3+/Fe2+ vs. Mo6+/Mo5+) can promote the formation of oxygen vacancies and enhance the activity of lattice oxygen, thereby improving surface stability [31]. To clearly justify our surface selection, we note that our stoichiometric surface-energy calculations show FeMoO- and FeO2-terminated (001) surfaces are nearly degenerate in stability, with γFeMoO = 0.09172 eV Å−2 and γFeO2 = 0.09171 eV Å−2 (differing only in the fifth decimal place and thus within the accuracy of our DFT + U approach). However, the FeMoO termination is particularly advantageous because it simultaneously exposes both Fe and Mo cations, providing dual redox-active sites directly relevant to CO2/H2O adsorption and activation. By contrast, the SrO termination is significantly less stable (γ ≈ 0.146 eV Å−2). Previous DFT + U thermodynamic analysis also indicates that Mo-containing (plane-Mo) terminations can be competitive in stability, whereas FeO2 termination is favored only under strongly oxidizing conditions [32]. Considering both comparable energetic stability under stoichiometric conditions and the direct availability of Fe and Mo active sites, the FeMoO surface was selected as the preferred cleavage plane for subsequent adsorption and activation studies.
We investigated both single and co-adsorption of CO2 and H2O molecules on the FeMoO-terminated surface. The configurations of CO2 adsorption are shown in Figure S2. CO2 molecules were initially placed with their molecular axis either perpendicular or parallel to the surface over Fe, Mo, and O surface sites. Configurations (a,b) correspond to adsorption on Fe sites with the molecular axis perpendicular (a) or parallel (b) to the surface, (c,d) to adsorption on Mo sites with similar perpendicular (c) and parallel (d) orientations, and adsorption on O sites was considered by allowing the carbon atom to interact directly with surface oxygen atoms. Upon optimization (Figure 3), CO2 adsorbed on metal sites tended to reorient to perpendicular configurations, whereas adsorption over O sites led to full activation of the CO2 molecule. Adsorption energies, bond lengths, and bond angles for each configuration are summarized in Table 2. In the table, “-1” denotes perpendicular adsorption (molecular axis roughly perpendicular to the surface), “-2” denotes parallel adsorption (molecular axis roughly parallel to the surface), and the preceding letter indicates the adsorption site. It should be noted that because the surface slab is periodic in the lateral directions, all top-layer Fe, Mo, and O atoms are crystallographically equivalent; therefore, sampling one representative Fe, Mo, and O site exhausts all non-equivalent adsorption centers on this surface.
While our comprehensive screening identified initial adsorption configurations across all geometrically plausible sites, thermodynamic analysis revealed fundamental limitations. For instance, CO2 adsorption on the Mo-1 site yielded a positive adsorption energy (Eads = +0.959 eV), which indicates that this configuration is thermodynamically unstable and would spontaneously desorb. Such unstable cases are therefore excluded from further consideration as viable intermediates. In stark contrast, among the thermodynamically stable configurations, CO2 adsorption at the O site exhibited exceptionally strong interaction, with Eads = −2.047 eV. This pronounced binding was accompanied by significant molecular distortion, including contraction of the C=O bond and alteration of the O-C-O bond angle, indicative of profound electronic interaction with the surface. The short adsorption distance and substantial electron transfer observed in the O-2 configuration further suggest enhanced CO2 activation, likely mediated through covalent character or significant polarization effects.
To further explore the electronic behavior of CO2 adsorption, we computed the charge density difference and Bader charges for the most favorable configurations: Fe-1, Mo-2, and O-2. The charge density difference plots are shown in Figure 4. Yellow regions indicate electron accumulation, while blue regions indicate depletion. An isosurface value of 0.002 e/Å3 was used. Minor electron transfer was observed between CO2 and the Fe or Mo atoms at the respective adsorption sites, supporting the presence of moderate chemical interaction. In contrast, the O-2 configuration displayed significant charge redistribution between the CO2 molecule and surface oxygen atoms, indicating a stronger interaction possibly involving covalent character.
Quantitative analysis using Bader charge (Table 3 and Figure S3) further confirms this behavior. Since the calculated actual number of charges has no absolute physical meaning, the charge state of each atom is determined by comparing its calculated charge with the corresponding valence charge, and the average Bader net atomic charge is obtained by averaging the charge transfer of all atoms of the same type. In this convention, a negative charge indicates that the atom has gained electrons, while a positive charge indicates electron loss relative to the isolated atom. The average net Bader charge on O atoms in the CO2 molecule for the O-2 configuration is −1.199 e, the most negative among all configurations, indicating substantial electron gain. This trend is consistent with the adsorption energy and charge density difference results.
To further clarify the surface redox behavior, we calculated the net charge transfer to the entire adsorbate molecule by summing the Bader charges of all its constituent atoms and comparing them with the isolated molecule. For O-site CO2 adsorption, the molecule gains a total of 0.53 e, with the C atom (+0.608 e) and the two O atoms bound to Mo (−0.629 e) and Fe (−0.513 e) serving as the main electron-accepting sites (Figure S3a). The adjacent Fe (1.653 e) and Mo (3.114 e) centers exhibit decreased electron density compared to the pristine surface, consistent with partial oxidation of these cations during CO2 activation.
Through the adsorption of a single CO2 molecule, it was found that CO2 preferentially adsorbs on the SFM surface with the carbon atom binding to surface oxygen sites. Fe and Mo atoms dynamically activate two adjacent oxygen atoms, which helps to partially weaken the strong C=O double bonds. This bond weakening facilitates subsequent surface reactions.
Similarly, the adsorption behavior of a single H2O molecule on the SFM surface was investigated, as shown in Figure S4. Panels (a,b) illustrate adsorption configurations at the Fe site with the molecular plane either roughly parallel to the surface or with one O–H bond pointing toward the surface, while panels (c,d) show the corresponding configurations at the Mo site. For adsorption at the O site, only interactions between hydrogen atoms and surface oxygen atoms were considered. The optimized structures (Figure 5) reveal that H2O molecules at metal sites tend to adopt configurations with the molecular plane parallel to the surface, which contrasts with the perpendicular adsorption observed for CO2. In the case of initial adsorption at the oxygen site, H2O molecules eventually migrate toward adjacent metal atoms and stabilize there. This suggests that H2O prefers to interact with metal sites on the surface, where hydrogen atoms oriented toward the surface are more likely to form stable hydrogen bonds with surface oxygen atoms.
Table 4 summarizes the adsorption energy, bond lengths, bond angles, and surface distances of H2O molecules at different adsorption sites. Since H2O molecules tend to interact with nearby metal atoms regardless of the initial placement configuration, the bond lengths, bond angles, and surface distances across different configurations do not vary significantly. Notably, the Mo-1 site exhibits the highest adsorption energy, reaching −0.900 eV. These results indicate that H2O adsorption can adopt multiple local configurations—ranging from chemisorption to physisorption—depending on whether the hydrogen atoms form hydrogen bonds with surface oxygen atoms, or even desorb under certain conditions.
To further investigate the electronic behavior of H2O adsorption, charge density difference and Bader charge analyses were performed. Given the unique adsorption characteristics of H2O, we focused on comparing the two most stable adsorption configurations at metal sites—Fe-2 and Mo-1. The isosurface value for the charge difference plots in Figure 6 was set to 0.002. It is evident from Figure 6 that there is significant charge accumulation between the surface metal atoms and the oxygen atom in the H2O molecule, which is likely a key factor contributing to stable adsorption on the SFM surface. The Bader charge analysis in Table 5 shows minimal differences in charge transfer at the metal sites between the two configurations, with only slight variations observed in the electron gain of the hydrogen atoms. Combined with the similar adsorption energies in both cases, these results indicate that H2O molecules can stably adsorb with their molecular plane roughly parallel to the surface at both Fe and Mo sites on the SFM surface.
To quantify the electron redistribution, the net charge transfer to the entire H2O molecule was evaluated (Figure S3b). For adsorption on a Mo site, the molecule undergoes a net electron loss of 0.11 e (O: +0.641 e; H: −0.400 e, −0.355 e), while the coordinated Mo atom becomes slightly oxidized (3.141 e), indicating electron donation from water to the surface.
Building on the results of individual CO2 and H2O adsorption, we investigated the co-adsorption behavior of CO2 and H2O on the SFM surface. Among several possible co-adsorption geometries, the configuration shown in Figure 7—H2O adsorbed on the Mo site with CO2 nearby—was found to be the most stable. As shown in Figure 7, the H2O molecule remains adsorbed with its molecular plane roughly parallel to the Mo site, consistent with the single adsorption results. However, the CO2 molecule, influenced by the presence of H2O at the interface, adopts a configuration with its molecular axis roughly parallel to the surface. The corresponding charge density difference map in Figure 7b reveals that CO2 adsorption significantly alters the surface electronic structure by inducing pronounced charge redistribution at metal sites, thereby affecting the overall co-adsorption process.
Furthermore, we calculated the co-adsorption energy and differential adsorption energies to evaluate the interaction between CO2 and H2O. As shown in Figure 8, the co-adsorption energy is −1.378 eV, indicating that the simultaneous adsorption of CO2 and H2O on the surface is energetically favorable, although the stabilization is less pronounced compared to single-molecule adsorption. Additionally, the charge density difference map (Figure 7) reveals substantial electron redistribution between the adsorbates, confirming the presence of intermolecular interactions. The differential adsorption energy analysis shows that when CO2 adsorbs first, the subsequent H2O adsorption is unfavorable, with a differential adsorption energy of +0.670 eV, indicating that pre-adsorbed CO2 can hinder H2O binding at the same site. In contrast, when H2O is adsorbed first, the subsequent CO2 adsorption remains favorable, with a differential adsorption energy of −0.477 eV. These results suggest that, during surface co-electrolysis processes on SFM, initial H2O adsorption does not prevent CO2 from binding, whereas pre-adsorbed CO2 may block H2O adsorption sites, potentially affecting subsequent surface reactions involving water.
To identify the electronic states responsible for CO2 and H2O adsorption, we analyzed the projected density of states (PDOS) for surface Fe-3d, Mo-3d, O-2p, as well as molecular C-2p, C-1s, and H-1s states (Figure 9 and Figures S5–S7). For CO2 adsorbed at a surface O site, the C-2p states show modest overlap with surface O-2p states in the −8 to −6 eV range, with the peak near −7 eV also intersecting weakly with Fe-3d and Mo-3d states. This indicates σ-bond formation and partial π* back-donation from surface O to the CO2 antibonding orbitals, consistent with moderate electron transfer to CO2, as reflected in the differential charge analysis [33]. In the 4–5 eV region, C-2p again overlaps with O-2p, Fe-3d, and Mo-3d, with spin-down Fe-3d exhibiting slightly higher intensity, highlighting the role of Fe d-orbitals in mediating surface reactivity and stabilizing the activated CO2 [34].
For H2O adsorption at a Mo site, H-1s states overlap with O-2p states near −10 eV, confirming O–H bond formation. Fe-3d states near the Fermi level show similar features to CO2 adsorption, while Mo-3d contributions are slightly diminished, indicating that Fe is more responsive to adsorbate interaction than Mo [35].
In the co-adsorption configuration, C-2p states around 3 eV overlap with Fe-3d, Mo-3d, and O-2p states. Spin-up C-2p shows the largest intensity, followed by O-2p and Mo-3d, whereas spin-down Fe-3d dominates relative to other orbitals, reflecting competition for Fe-centered states. These observations, together with differential charge, suggest that CO2 and H2O interact synergistically via surface O atoms, with Fe-3d orbitals being highly sensitive to adsorption environment and mediating electron redistribution, while Mo-3d states serve a stabilizing role [36].
In addition, bader analysis (Figure S3c and Table S1) shows ΔQ(CO2) ≈ −0.001 e (C: +1.607 e; O: −0.774 e, −0.834 e), indicating no appreciable net charge transfer to CO2 but pronounced intramolecular polarization (electron accumulation on O and depletion on C). Water gains a small amount of charge, ΔQ(H2O) = −0.047 e (H: +0.303 e, +0.382 e; O: −0.732 e). The adjacent cations become more reduced relative to the single-adsorption cases, with Fe = 1.567 e and Mo = 2.325 e, consistent with enhanced Fe/Mo–O covalency rather than large net electron flow to the adsorbates.
The sign reversal of the Bader charge on the water O atom from +0.641 e (single adsorption on Mo) to −0.732 e (co-adsorption) can be rationalized by a change in adsorption geometry, electronic coupling mode, and local electrostatic environment. In the single-adsorption case, H2O primarily binds via σ-donation from the O lone pair into Mo-3d states, which depletes electron density on O (positive Bader charge) and renders the H atoms slightly electron-rich. In the co-adsorbed configuration, the strongly polarized CO2 (electron-rich O, electron-deficient C) and the resulting H···O(CO2) hydrogen bonds reorient H2O and increase the degree of metal–oxygen covalency. Consistently, the two H atoms become more positive (H: +0.303/+0.382 e) while the O atom accumulates electron density (O: −0.732 e). This is in line with the PDOS, which exhibits a modest increase in Mo-3d intensity near E_F in the co-adsorption state, suggestive of strengthened d–p hybridization between Mo-3d and O-2p orbitals.
The sign reversal of the Bader charge on the water O atom from +0.641 e (single adsorption on Mo) to −0.732 e (co-adsorption) can be rationalized by a change in adsorption geometry, electronic coupling mode, and local electrostatic environment. In the single-adsorption case, H2O primarily binds via σ-donation from the O lone pair into Mo-3d states, which depletes electron density on O (positive Bader charge) and renders the H atoms slightly electron-rich. In the co-adsorbed configuration, the strongly polarized CO2 (electron-rich O, electron-deficient C) and the resulting H···O(CO2) hydrogen bonds reorient H2O and increase the degree of metal–oxygen covalency. Consistently, the two H atoms become more positive (H: +0.303/+0.382 e) while the O atom accumulates electron density (O: −0.732 e). This is in line with the PDOS, which exhibits a modest increase in Mo-3d intensity near Ef in the co-adsorption state, suggestive of strengthened d–p hybridization between Mo-3d and O-2p orbitals.
Taken together with the PDOS results, these findings indicate that co-adsorption primarily polarizes CO2 and H2O and strengthens metal–oxygen bonding, which could lower the barriers for subsequent proton–electron-coupled steps, even in the absence of significant net charge transfer to CO2. These electronic features have direct implications for electrochemical co-electrolysis mechanisms. For CO2, electron accumulation on the O atoms and depletion at the C center indicate weakening of the C–O bonds, which is typically associated with the CO pathway. At the same time, polarization of the C atom suggests that proton attack could also stabilize a formate (HCOO) intermediate, consistent with competing reduction routes reported in experiments. For H2O, the hybridization between H-1s and O-2p orbitals and the observed charge accumulation around the O atom point to facilitated O–H bond cleavage, a critical step for providing protons and hydroxyl species during co-electrolysis. Together, these trends suggest that the Fe/Mo–O electronic states not only stabilize adsorption but also mediate the initial bond-breaking steps required for syngas evolution. Importantly, these microscopic insights align with experimental SOEC studies, which report that SFM-based cathodes exhibit relatively low polarization resistance and moderate activation energies [37,38]. Such macroscopic performance indicators are consistent with our finding that Fe/Mo–O hybridized states enhance the ease of CO2 activation and H2O dissociation. While our DFT calculations do not directly predict electrochemical resistances, the identified electronic interactions provide a mechanistic rationale for the favorable kinetics observed experimentally, highlighting the connection between atomic-scale electronic structure and operando catalytic behavior.
It should be emphasized, however, that the present study was performed on a defect-free SFM surface. In real SOEC environments, oxygen vacancies are prevalent and have been shown to significantly influence adsorption energetics and redox chemistry [23]. Therefore, the present results should be regarded as a theoretical benchmark, with future work extending to defective surfaces to better capture realistic catalytic conditions. Beyond the static adsorption configurations considered here, the interplay between CO2 and H2O is expected to be strongly dependent on surface coverage and operating environment. At moderate coverage, pre-adsorbed H2O can stabilize CO2 via hydrogen bonding and enhanced Fe/Mo–O covalency, leading to a synergistic effect that may lower activation barriers for proton–electron transfer. In contrast, excessive CO2 coverage may inhibit H2O adsorption by occupying adjacent oxygen sites, thereby reducing the availability of surface protons and suppressing co-electrolysis activity. Under operando conditions, elevated temperatures (700–900 °C) are likely to shift adsorption–desorption equilibria toward dynamic exchange, while applied cathodic bias can enhance charge transfer to CO2, favoring its reduction pathway. These considerations suggest that the balance between synergy and inhibition is not fixed but instead depends sensitively on both surface coverage and reaction environment, providing guidance for interpreting experimental trends and designing more robust SOEC cathodes.

3. Computational Approach

All calculations were performed within the framework of density functional theory (DFT) using the projector-augmented wave (PAW) method, as implemented in the Vienna Ab initio Simulation Package (VASP 5.4.4) [39]. The exchange-correlation effects were described using the Perdew-Burke-Ernzerhof (PBE) functional within the generalized gradient approximation (GGA) [40]. A plane-wave energy cutoff of 520 eV was employed. Brillouin zone sampling was carried out using the Monkhorst-Pack scheme, with a k-point grid of 2 × 2 × 1 for surface calculations and 4 × 4 × 3 for bulk calculations, ensuring adequate convergence. The electronic self-consistent field (SCF) convergence criterion was set to 10−6 eV. Structural optimizations were performed until the residual forces on all atoms were reduced to below 0.05 eV/Å. A vacuum layer of 15 Å was applied along the surface normal to eliminate spurious interactions between periodic images. To account for long-range dispersion interactions, the Grimme DFT-D2 method [41] was employed in all calculations. Since GGA-based DFT tends to underestimate band gaps due to self-interaction errors, on-site Coulomb interactions were incorporated using the Hubbard U approach. Specifically, U values of 4.0 eV and 3.3 eV were applied to the Fe 3d and Mo 3d orbitals, respectively. These values are widely used in previous studies and related Fe/Mo-based perovskites, and have been shown to reproduce structural, electronic, and catalytic properties with good accuracy [42,43,44]. Additionally, spin polarization was considered in all calculations.
In terms of adsorption energy, the formula is as follows:
E ads   =   E gas   +   surface     E gas     E surface
where E gas   +   surface and E gas are the calculated energies of the model after adsorption of the gas on the surface and the free gas molecule, respectively. E surface is the total energy of the relaxed clean surface.
The differential adsorption energy was calculated to assess the effect of pre-adsorbed molecule 1 on the adsorption of molecule 2:
E ads , m o l 2 / m o l 1   =   E surface   +   m o l 1   +   m o l 2     E surface   +   m o l 1     E m o l 2
A decrease in the adsorption energy compared to the isolated adsorption case suggests that molecule 1 promotes the adsorption of molecule 2, likely through induced dipole interactions or charge transfer. Conversely, a reduced adsorption affinity may indicate a competitive adsorption mechanism, where molecule 1 modifies the local electronic structure of the surface, making it less favorable for molecule 2 adsorption.
The formula for calculating co-adsorption energy is as follows:
E c o - a d s   =   E surface   +   m o l 1   +   m o l 2     E surface     E m o l 1     E m o l 2
Charge Density Difference refers to the change in the density of electrons in a certain state (e.g., molecule formation, adsorption process) relative to the initial state (e.g., isolated atoms). The formula is usually expressed as:
ρ   =   ρ Total     ( ρ surface   +   ρ mol )
where ρ Total is the electron density of the combined adsorbate-surface system, ρ surface and ρ mol are the electron densities of the isolated surface and molecule, respectively. Both isolated densities are computed with the atomic positions fixed to those in the combined system to accurately capture charge rearrangement due to adsorption.
For symmetric slabs the surface free energy γ was evaluated as:
γ   =   E slab       N fu E bulk     N O μ O 2 A
where E slab is the total energy of the slab, E bulk is the energy per bulk formula unit, N fu is the number of formula units in the slab, N O is the oxygen excess/deficit relative to bulk stoichiometry, and A is the surface area of one side. For our slab constructions N O = 0 (each slab contains 72 O atoms while the bulk formula unit contains 12 O atoms, hence N fu = 6), so N O   = 0 and γ reduces to ( E slab     N fu E bulk )/(2A). Using E bulk   = −138.572 eV and A = 184.265 Å2.
To quantitatively assign the redistributed charge to individual atoms, we employed Bader charge population analysis, which partitions the electron density using real-space zero-flux surfaces in ∇ρ. Such a topology-based scheme is particularly suitable for plane-wave DFT calculations, as it avoids the basis-set dependence inherent in Mulliken- or Löwdin-type population analyses that are commonly used in localized (Gaussian) basis sets. All Bader charge values reported in this work are expressed in units of the elementary charge |e|.

4. Conclusions

In summary, we conducted a comprehensive DFT investigation on the adsorption behavior of CO2 and H2O molecules on the FeMoO-terminated (001) surface of Sr2FeMoO6 (SFM) perovskite. The results show that CO2 strongly interacts with surface oxygen sites with a binding energy of −2.047 eV, leading to notable molecular activation and substantial charge transfer, while H2O displays weaker adsorption (−0.900 eV) and preferentially binds to Mo sites. Charge density difference and Bader charge analyses confirm the distinct electronic responses of the surface upon adsorption, revealing covalent bonding features in CO2 adsorption and more electrostatic characteristics in H2O interaction. Co-adsorption studies indicate that the presence of H2O modifies the adsorption geometry of CO2 and enhances the co-adsorption energy of CO2 and H2O on the surface, suggesting a synergistic effect that may facilitate surface proton–electron coupling processes. However, pre-adsorbed CO2 weakens the adsorption of H2O, and due to competitive site occupation, excessive CO2 coverage could inhibit the activation of H2O. Density of states analysis suggests that the Fe/Mo–O hybridized states near the Fermi level play a crucial role in mediating surface reactivity. It should be noted that all results were obtained on a perfect, defect-free SFM surface, which, while idealized compared to real materials containing vacancies or mixed-valence cations, serves as a theoretical benchmark for understanding fundamental molecule–surface and molecule–molecule interactions. Overall, these findings contribute to a molecular-level understanding of the surface interaction mechanisms of perovskite materials under conditions relevant to CO2 + H2O co-electrolysis and offer theoretical guidance for the rational design of high-performance double perovskite electrocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15090884/s1, Figure S1: Optimized geometries for (a) FeMoO (001) surface, (b) FeO2 (001) surface, and (c) SrO (001) surface; Figure S2: The initial adsorption and optimal configurations of CO2 on the surface; (a–e) are the initial adsorption configurations, and (f–j) are the optimized configurations; Figure S3: Optimized adsorption geometries with atom-resolved Bader charges for (a) CO2 adsorption on O site, (b) H2O adsorption on Mo-2 site, and (c) CO2 + H2O co-adsorption on the SFM surface; Figure S4: The initial adsorption configuration and optimal configuration of H2O on the surface; (a–e) are the initial adsorption configurations, and (f–j) are the optimized configurations; Figure S5: PDOS of O-CO2; Figure S6: PDOS of Mo-H2O; Figure S7: PDOS of CO2+H2O; Table S1: The calculated Bader charge for O-CO2, Mo-H2O, and CO2+H2O model.

Author Contributions

Conceptualization, J.W. and Q.Z.; methodology, Z.Z., S.H., N.L. and Y.G.; validation, K.D., Z.Z., S.H., N.L. and Y.G.; resources, Y.W., J.Z. and K.W.; writing—original draft preparation, J.W. and Q.Z.; writing—review and editing, J.W.; visualization, Z.Z., S.H., N.L., Y.G. and K.D.; supervision, Y.W., J.Z. and K.W.; funding acquisition, Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Project of China Southern Power Grid Co., Ltd. (Grant No. GDKJXM20231333), entitled “Key Technology for Power-to-Gas Based on Coupled Co-electrolysis of CO2 and H2O”.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors thank the Science and Technology Project of China Southern Power Grid Co., Ltd. (Grant No. GDKJXM20231333), entitled “Key Technology for Power-to-Gas Based on Coupled Co-electrolysis of CO2 and H2O”.

Conflicts of Interest

Authors Junbo Wang, Zixuan Zhang, Sijie He and Nianbo Liang was employed by the company Foshan Power Supply Bureau of Guangdong Power Grid. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. The authors declare that this study received funding from Science and Technology Project of China Southern Power Grid Co., Ltd. The funder was not involved in the study design, collection, analysis, interpretation of data, the writing of this article or the decision to submit it for publication.

References

  1. Deka, D.J.; Kim, J.; Gunduz, S.; Ferree, M.; Co, A.C.; Ozkan, Ü.S. Temperature-induced changes in the synthesis gas composition in a high-temperature H2O and CO2 co-electrolysis system. Appl. Catal. A Gen. 2020, 602, 117697. [Google Scholar] [CrossRef]
  2. Götz, M.; Lefebvre, J.; Mörs, F.; Koch, A.M.; Graf, F.; Bajohr, S.; Reimert, R.; Kolb, T. Renewable Power-to-Gas: A technological and economic review. Renew. Energy 2016, 85, 1371–1390. [Google Scholar] [CrossRef]
  3. Zheng, Y.; Wang, J.; Yu, B.; Zhang, W.; Chen, J.; Qiao, J.; Zhang, J. A review of high temperature co-electrolysis of H2O and CO2 to produce sustainable fuels using solid oxide electrolysis cells (SOECs): Advanced materials and technology. Chem. Soc. Rev. 2017, 46, 1427–1463. [Google Scholar] [CrossRef] [PubMed]
  4. Machado, M.; de Souza Junior, R.L.; de Almeida, J.M.A.R.; Romano, P.N.; Garcia, M.A.S. Syngas enhancement for Fischer-Tropsch integration via solid oxide electrolyzer cell co-electrolysis with or without methane. iScience 2024, 27, 111014. [Google Scholar] [CrossRef]
  5. Gao, S.; Lin, Y.; Jiao, X.; Sun, Y.; Luo, Q.; Zhang, W.; Li, D.; Yang, J.; Xie, Y. Partially oxidized atomic cobalt layers for carbon dioxide electroreduction to liquid fuel. Nature 2016, 529, 68–71. [Google Scholar] [CrossRef] [PubMed]
  6. Zanatta, M. Materials for Direct Air Capture and Integrated CO2 Conversion: Advancement, Challenges, and Prospects. ACS Mater. Au 2023, 3, 576–583. [Google Scholar] [CrossRef]
  7. Adamu, A.; Russo-Abegão, F.; Boodhoo, K. Process intensification technologies for CO2 capture and conversion—A review. BMC Chem. Eng. 2020, 2, 2. [Google Scholar] [CrossRef]
  8. Zanatta, M.; García-Verdugo, E.; Sans, V. Direct Air Capture and Integrated Conversion of Carbon Dioxide into Cyclic Carbonates with Basic Organic Salts. ACS Sustain. Chem. Eng. 2023, 11, 9613–9619. [Google Scholar] [CrossRef]
  9. Yang, L.; Cheng, Z.; Liu, M.; Wilson, L. New insights into sulfur poisoning behavior of Ni-YSZ anode from long-term operation of anode-supported SOFCs. Energy Environ. Sci. 2010, 3, 1804–1809. [Google Scholar] [CrossRef]
  10. Liao, Y.; Xi, X.; Chen, H.; Liu, J.; Fu, X.Z.; Luo, J.L. The emerging Sr2FeMoO6-based electrocatalysts for solid oxide electrochemical cell: Synthesis, modulation and applications. Chem. Synth. 2024, 4, 18. [Google Scholar] [CrossRef]
  11. Bibi, B.; Nazar, A.; Zhu, B.; Yang, F.; Yousaf, M.; Raza, R.; Shah, M.A.K.Y.; Kim, J.S.; Afzal, M.; Lei, Y.; et al. Emerging semiconductor ionic materials tailored by mixed ionic-electronic conductors for advanced fuel cells. Adv. Powder Mater. 2024, 3, 100231. [Google Scholar] [CrossRef]
  12. Han, Z.; Wang, Y.; Yang, Y.; Li, L.; Yang, Z.; Han, M. High-performance SOFCs with impregnated Sr2Fe1.5Mo0.5O6−δ anodes toward sulfur resistance. J. Alloys Compd. 2017, 703, 258–263. [Google Scholar] [CrossRef]
  13. Ma, Y.; Zhang, L.; Zhu, K. In3+-doped Sr2Fe1.5Mo0.5O6−δ cathode with improved performance for an intermediate-temperature solid oxide fuel cell. Nano Res. 2023, 17, 407–415. [Google Scholar] [CrossRef]
  14. Zhang, L.; Yin, Y.; Xu, Y.; Yu, S.; Bi, L. Tailoring Sr2Fe1.5Mo0.5O6−δ with Sc as a new single-phase cathode for proton-conducting solid oxide fuel cells. Sci. China Mater. 2022, 65, 1485–1494. [Google Scholar] [CrossRef]
  15. Muñoz-García, A.B.; Pavone, M.; Ritzmann, A.M.; Carter, E.A. Oxide ion transport in Sr2Fe1.5Mo0.5O6−δ, a mixed ion-electron conductor: New insights from first principles modeling. Phys. Chem. Chem. Phys. 2013, 15, 6250–6259. [Google Scholar] [CrossRef]
  16. Gou, M.L.; Ren, R.Z.; Sun, W.; Xu, C.M.; Meng, X.G.; Wang, Z.H.; Qiao, J.S.; Sun, K.N. Nb-doped Sr2Fe1.5Mo0.5O6−δ electrode with enhanced stability and electrochemical performance for symmetrical solid oxide fuel cells. Ceram. Int. 2019, 45, 15696–15704. [Google Scholar] [CrossRef]
  17. Guhl, H.; Miller, W.; Reuter, K. Water adsorption and dissociation on SrTiO3 (001) revisited: A density functional theory study. Phys. Rev. B 2010, 81, 155455. [Google Scholar] [CrossRef]
  18. Baniecki, J.D.; Ishii, M.; Kurihara, K.; Yamanaka, K.; Yano, T.; Shinozaki, K.; Imada, T.; Kobayashi, Y. Chemisorption of water and carbon dioxide on nanostructured BaTiO3–SrTiO3 (001) surfaces. J. Appl. Phys. 2009, 106, 054109. [Google Scholar] [CrossRef]
  19. Wang, Z.; Hao, X.; Gerhold, S.; Novotny, Z.; Franchini, C.; McDermott, E.; Schulte, K.; Schmid, M.; Diebold, U. Water Adsorption at the Tetrahedral Titania Surface Layer of SrTiO3 (110)-(4 × 1). J. Phys. Chem. C 2013, 117, 26060–26069. [Google Scholar] [CrossRef] [PubMed]
  20. Becerra-Toledo, A.E.; Enterkin, J.A.; Kienzle, D.M.; Marks, L.D. Water adsorption on SrTiO3 (001): II. Water, water, everywhere. Surf. Sci. 2012, 606, 791–802. [Google Scholar] [CrossRef]
  21. Reuter, K.; Scheffler, M. Composition, structure, and stability of RuO2 (110) as a function of oxygen pressure. Phys. Rev. B 2001, 65, 035406. [Google Scholar] [CrossRef]
  22. Rojas, E.; Calatayud, M.; Bañares, M.A.; Guerrero-Pérez, M.O. Theoretical and Experimental Study of Light Hydrocarbon Ammoxidation and Oxidative Dehydrogenation on (110)-VSbO4 Surfaces. J. Phys. Chem. C 2012, 116, 9132–9141. [Google Scholar] [CrossRef]
  23. Sun, Y.; Zhou, J.; Yang, J.; Neagu, D.; Liu, Z.; Yin, C.; Xue, Z.; Zhou, Z.; Cui, J.; Wu, K. Nanosurface-Reconstructed Fuel Electrode by Selective Etching for Highly Efficient and Stable Solid Oxide Cells. Adv. Sci. 2024, 12, 2409272. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, Q.; Wang, Y.; Jia, Y.; Yan, W.; Li, Q.; Zhou, J.; Wu, K. Engineering the electronic structure towards visible lights photocatalysis of CaTiO3 perovskites by cation (La/Ce)-anion (N/S) co-doping: A first-principles study. Molecules 2023, 28, 7134. [Google Scholar] [CrossRef]
  25. Wang, Y.; Zhou, Q.; Zhang, Q.; Ren, Y.; Cui, K.; Cheng, C.; Wu, K. Effects of La-N Co-Doping of BaTiO3 on Its Electron-Optical Properties for Photocatalysis: A DFT Study. Molecules 2024, 29, 2250. [Google Scholar] [CrossRef]
  26. Yang, M.; Yao, Z.; Liu, S.; Wang, J.; Sun, A.; Xu, H.; Yang, G.; Ran, R.; Zhou, W.; Xiao, G.; et al. Bismuth doped Sr2Fe1.5Mo0.5O6−δ double perovskite as a robust fuel electrode in ceramic oxide cells for direct CO2 electrolysis. J. Mater. Sci. Technol. 2023, 164, 160–167. [Google Scholar] [CrossRef]
  27. Wang, Z.; Tan, T.; Du, K.; Zhang, Q.; Liu, M.; Yang, C. A high-entropy layered perovskite coated with in situ exsolved core-shell CuFe@FeOx nanoparticles for efficient CO2 electrolysis. Adv. Mater. 2023, 36, 2312119. [Google Scholar] [CrossRef] [PubMed]
  28. Retuerto, M.; Li, M.R.; Go, Y.B.; Ignatov, A.; Croft, M.; Ramanujachary, K.V.; Hadermann, J.; Hodges, J.P.; Herber, R.H.; Nowik, I.; et al. Magnetic and Structural Studies of the Multifunctional Material SrFe0.75Mo0.25O3−δ. Inorg. Chem. 2012, 51, 12273–12282. [Google Scholar] [CrossRef]
  29. Bugaris, D.E.; Hodges, J.P.; Huq, A.; Chance, W.M.; Heyden, A.; Chen, F.; zur Loye, H.C. Investigation of the high-temperature redox chemistry of Sr2Fe1.5Mo0.5O6−δ via in situ neutron diffraction. J. Mater. Chem. A 2014, 2, 4045–4054. [Google Scholar] [CrossRef]
  30. Osinkin, D.A.; Beresnev, S.M.; Khodimchuk, A.V. Functional properties and electrochemical performance of Ca-doped Sr2−xCaxFe1.5Mo0.5O6−δ as anode for solid oxide fuel cells. J. Solid State Electrochem. 2018, 23, 627–634. [Google Scholar] [CrossRef]
  31. Sun, H.; Xu, X.; Song, Y.; Shao, Z. Recent progress in Sr2Fe1.5Mo0.5O6−δ-based multifunctional materials for energy conversion and storage. Adv. Funct. Mater. 2024, 14, 2411622. [Google Scholar] [CrossRef]
  32. Suthirakun, S.; Ammal, S.C.; Muñoz-García, A.B.; Xiao, G.; Chen, F.; zur Loye, H.-C.; Carter, E.A.; Heyden, A. Theoretical Investigation of H2 Oxidation on the Sr2Fe1.5Mo0.5O6 (001) Perovskite Surface under Anodic Solid Oxide Fuel Cell Conditions. J. Am. Chem. Soc. 2014, 136, 8374–8386. [Google Scholar] [CrossRef] [PubMed]
  33. Cheng, Z.; Sherman, B.J.; Lo, C.S. Carbon dioxide activation and dissociation on ceria (110): A density functional theory study. J. Chem. Phys. 2013, 138, 014702. [Google Scholar] [CrossRef] [PubMed]
  34. Pavelec, J.; Hulva, J.; Halwidl, D.; Bliem, R.; Gamba, O.; Jakub, Z.; Brunbauer, F.; Schmid, M.; Diebold, U.; Parkinson, G.S. A multi-technique study of CO2 adsorption on Fe3O4 magnetite. J. Chem. Phys. 2017, 146, 014701. [Google Scholar] [CrossRef]
  35. Hu, J.; Wang, C.; He, S.; Zhu, J.; Wei, L.; Zheng, S. A DFT-Based Model on the Adsorption Behavior of H2O, H⁺, Cl, and OH on Clean and Cr-Doped Fe(110) Planes. Coatings 2018, 8, 51. [Google Scholar] [CrossRef]
  36. Zeng, X.; Zeng, B.; Huang, L.; Zhong, L.; Li, X.; Huang, W. Adsorption of Y(III) on the Interface of Kaolinite-H2O: A DFT Study. Minerals 2022, 12, 1128. [Google Scholar] [CrossRef]
  37. Liu, Q.; Yang, C.; Dong, X.; Chen, F. Perovskite Sr2Fe1.5Mo0.5O6−δ as electrode materials for symmetrical solid oxide electrolysis cells. Int. J. Hydrogen Energy 2010, 35, 10039–10044. [Google Scholar] [CrossRef]
  38. Jiang, Y.; Yang, Y.; Xia, C.; Bouwmeester, H.J.M. Sr2Fe1.4Mn0.1Mo0.5O6−δ perovskite cathode for highly efficient CO2 electrolysis. J. Mater. Chem. A 2019, 7, 22939–22949. [Google Scholar] [CrossRef]
  39. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  40. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  41. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  42. Setyawan, W.; Curtarolo, S. High-throughput electronic band structure calculations: Challenges and tools. Comp. Mater. Sci. 2010, 49, 299–312. [Google Scholar] [CrossRef]
  43. Han, Z.; Dong, H.; Wu, Y.; Yang, Y. Locating the rate-limiting step of hydrogen conversion on Sr2Fe1.5Mo0.5O6 (001) surface: Implications for efficient SOFC anode design. Appl. Surf. Sci. 2022, 595, 153513. [Google Scholar] [CrossRef]
  44. Muñoz-García, A.B.; Pavone, M.; Carter, E.A. Effect of Antisite Defects on the Formation of Oxygen Vacancies in Sr2FeMoO6: Implications for Ion and Electron Transport. Chem. Mater. 2011, 23, 4525–4536. [Google Scholar] [CrossRef]
Figure 1. Optimized crystal structure.
Figure 1. Optimized crystal structure.
Catalysts 15 00884 g001
Figure 2. Optimized crystal structure (a) (001) Three views of the surface; (b) Oblique view.
Figure 2. Optimized crystal structure (a) (001) Three views of the surface; (b) Oblique view.
Catalysts 15 00884 g002
Figure 3. The optimal configurations of CO2 on the surface; (a) Fe-1-CO2; (b) Fe-2-CO2; (c) Mo-1-CO2; (d) Mo-2-CO2; (e) O-CO2.
Figure 3. The optimal configurations of CO2 on the surface; (a) Fe-1-CO2; (b) Fe-2-CO2; (c) Mo-1-CO2; (d) Mo-2-CO2; (e) O-CO2.
Catalysts 15 00884 g003
Figure 4. Charge Density Difference: (a) Fe-1-CO2; (b) Mo-2-CO2; (c) O-CO2.
Figure 4. Charge Density Difference: (a) Fe-1-CO2; (b) Mo-2-CO2; (c) O-CO2.
Catalysts 15 00884 g004
Figure 5. The optimal configurations of H2O on the surface; (a) Fe-1-H2O; (b) Fe-2-H2O; (c) Mo-1-H2O; (d) Mo-2-H2O; (e) O-H2O.
Figure 5. The optimal configurations of H2O on the surface; (a) Fe-1-H2O; (b) Fe-2-H2O; (c) Mo-1-H2O; (d) Mo-2-H2O; (e) O-H2O.
Catalysts 15 00884 g005
Figure 6. Charge Density Difference: (a) Fe-2-H2O; (b) Mo-1-H2O.
Figure 6. Charge Density Difference: (a) Fe-2-H2O; (b) Mo-1-H2O.
Catalysts 15 00884 g006
Figure 7. (a) Co-adsorption optimized structure; (b) differential charge.
Figure 7. (a) Co-adsorption optimized structure; (b) differential charge.
Catalysts 15 00884 g007
Figure 8. Comparison of CO2 and H2O co-adsorption energy.
Figure 8. Comparison of CO2 and H2O co-adsorption energy.
Catalysts 15 00884 g008
Figure 9. DOS results: (a) unadsorbed; (b) O-CO2; (c) Mo-H2O; (d) CO2 + H2O.
Figure 9. DOS results: (a) unadsorbed; (b) O-CO2; (c) Mo-H2O; (d) CO2 + H2O.
Catalysts 15 00884 g009
Table 1. Comparison of lattice parameters for Sr2Fe1.5Mo0.5O6 and related perovskite materials.
Table 1. Comparison of lattice parameters for Sr2Fe1.5Mo0.5O6 and related perovskite materials.
CompoundLattice Parameters (Å)MethodSpace GroupReference
Sr2Fe1.5Mo0.5O6a = 5.542 Å, c = 7.940 ÅGGA + UI4/mmmThis work
Sr2Fe1.5Mo0.5O6−δa ≈ 5.543 Å, c = 7.847 Åsol–gel synthesisI4/mmm [28]
Sr2Fe0.5Mo0.5O6a = 5.555 Å, c = 7.878 ÅIn situ powder neutron diffractionI4/mmm[29]
Sr2−xCaxFe1.5Mo0.5O6−δa = 5.55 Å, c = 7.90 ÅXRD (wet H2)I4/mmm[30]
Table 2. The adsorption energy, bond length, bond angle and distance from the surface of the CO2 adsorption configuration of SFM (d (Å)).
Table 2. The adsorption energy, bond length, bond angle and distance from the surface of the CO2 adsorption configuration of SFM (d (Å)).
Adsorption ConfigurationFe-1Fe-2Mo-1Mo-2O
Eads (eV)−1.853−1.7560.959 (unstable)−1.727−2.047
bond length (Å)1.1731.1781.1741.1791.256
bond angle (°)179.781178.822179.259178.567128.053
d (Å)2.4602.6792.5452.7411.343
Table 3. Bader charge.
Table 3. Bader charge.
Adsorption ConfigurationAverage Bader Net Atomic Charge (e)
SrFeMoO-SlabCO-CO2
Fe-11.5761.7022.644−1.1711.530−0.769
Mo-21.5741.6922.367−1.1451.860−0.942
O site1.6401.6043.095−1.1990.608−0.571
Table 4. The adsorption energy, bond length, bond angle, and distance from the surface of the SFM adsorption H2O configuration (d (Å)).
Table 4. The adsorption energy, bond length, bond angle, and distance from the surface of the SFM adsorption H2O configuration (d (Å)).
Adsorption ConfigurationFe-1Fe-2Mo-1Mo-2O
Eads (eV)1.396−0.816−0.9000.066−0.211
bond length (Å)0.9870.9840.9760.9770.978
bond angle (°)104.425103.107105.622105.776105.334
d (Å)1.9701.9852.3442.3472.345
Table 5. Bader charge.
Table 5. Bader charge.
Adsorption ConfigurationAverage Bader Net Atomic Charge (e)
SrFeMoO-SlabHO-H2O
Fe-11.6371.6193.082−1.205−0.4250.652
Mo-21.6371.6153.083−1.205−0.3780.641
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, J.; Zhang, Q.; Zhang, Z.; He, S.; Liang, N.; Gao, Y.; Deng, K.; Wang, Y.; Zhou, J.; Wu, K. Atomic-Scale Insights into CO2 and H2O Co-Adsorption on Sr2Fe1.5Mo0.5O6 Surfaces: Role of Electronic Structure and Dual-Site Interactions. Catalysts 2025, 15, 884. https://doi.org/10.3390/catal15090884

AMA Style

Wang J, Zhang Q, Zhang Z, He S, Liang N, Gao Y, Deng K, Wang Y, Zhou J, Wu K. Atomic-Scale Insights into CO2 and H2O Co-Adsorption on Sr2Fe1.5Mo0.5O6 Surfaces: Role of Electronic Structure and Dual-Site Interactions. Catalysts. 2025; 15(9):884. https://doi.org/10.3390/catal15090884

Chicago/Turabian Style

Wang, Junbo, Qiankai Zhang, Zixuan Zhang, Sijie He, Nianbo Liang, Yuan Gao, Ke Deng, Yang Wang, Jun Zhou, and Kai Wu. 2025. "Atomic-Scale Insights into CO2 and H2O Co-Adsorption on Sr2Fe1.5Mo0.5O6 Surfaces: Role of Electronic Structure and Dual-Site Interactions" Catalysts 15, no. 9: 884. https://doi.org/10.3390/catal15090884

APA Style

Wang, J., Zhang, Q., Zhang, Z., He, S., Liang, N., Gao, Y., Deng, K., Wang, Y., Zhou, J., & Wu, K. (2025). Atomic-Scale Insights into CO2 and H2O Co-Adsorption on Sr2Fe1.5Mo0.5O6 Surfaces: Role of Electronic Structure and Dual-Site Interactions. Catalysts, 15(9), 884. https://doi.org/10.3390/catal15090884

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