Next Article in Journal
Construction of BiOBr/BNQDs Heterostructure Photocatalyst and Performance Studies of Photocatalytic Degradation of RhB
Previous Article in Journal
Relationship Between Number and Strength of Acid–Base Catalytic Sites and Their Performances in Isopropanol Dehydration Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Study of Sn-Doped RuO2 as Efficient Electrocatalysts for Enhanced Oxygen Evolution

1
School of Physics, Nankai University, Tianjin 300071, China
2
Key Laboratory of Advanced Electrode Materials for Novel Solar Cells for Petroleum and Chemical Industry of China, School of Chemistry and Life Sciences, Suzhou University of Science and Technology, Suzhou 215009, China
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(8), 770; https://doi.org/10.3390/catal15080770
Submission received: 9 July 2025 / Revised: 7 August 2025 / Accepted: 11 August 2025 / Published: 13 August 2025

Abstract

Improving the catalytic performance of the oxygen evolution reaction (OER) for water splitting in acidic media is crucial for the production of clean and renewable hydrogen energy. Herein, we study the OER electrocatalytic properties of various active sites on four exposed (110) and ( 1 ¯ 10) surfaces of Sn-doped RuO2 (Sn/RuO2) with antiferromagnetic arrangements in acidic environments. The Sn/RuO2 bulk structure with the Cm space group exhibits favorable thermodynamic stability. The coordinatively unsaturated metal (Mcus) sites distributed on the right branch of the volcano plot are generally more active than the bridge-bonded lattice oxygen (Obr) sites located on the left. Different from the conventional knowledge that the most active site is located in the nearest neighbor of the doped atom, it has a lower OER overpotential when the active site is 3.6 Å away from the doped Sn atom. Among the sites studied, the 46-Rucus site exhibits the optimal OER catalytic performance. The inherent factors affecting the OER activity of each site on the Sn/RuO2 surface are further analyzed, including the center of the d/p band at the active sites, the average electrostatic potential of the ions, and the number of transferred electrons. This work provides a reminder for the selection of active sites used to evaluate catalytic performance, which will benefit the development of efficient OER electrocatalysts.

Graphical Abstract

1. Introduction

Since fossil fuel reserves are limited and their excessive use causes environmental pollution, it is urgent to explore alternative energy to replace fossil fuels [1,2,3]. Clean and renewable hydrogen (H2) energy generated by electrochemical water splitting is considered an ideal substitute and has received widespread attention [4,5,6]. The water splitting process involves two key reactions: the oxygen evolution reaction (OER) and the hydrogen evolution reaction (HER). Compared with HER, the slow kinetic OER greatly reduces the overall efficiency of H2 preparation [7,8,9]. In this case, researchers have been conducting in-depth studies on various OER catalysts to improve their catalytic performance, such as precious metals, transition metal oxides, single-atom catalysts, etc. [10,11,12,13,14] Especially in an acidic environment, the advantages of higher ionic conductivity, fewer unfavorable reactions, and faster system responses are beneficial to promote OER reaction rates [15,16,17,18]. Therefore, designing highly efficient OER catalysts in acidic conditions is of great significance for the development of sustainable energy.
Rutile-phase RuO2 with outstanding OER catalytic performance in acidic media has always been a benchmark catalyst [19,20,21]. Given that electrochemical reactions usually occur on the catalyst surface, the introduction of doped atoms is a common effective strategy to improve the efficiency of electrocatalysts, which will affect properties such as the number of exposed active sites, intrinsic activity, and electronic structure [22]. For example, Chen et al. indicated that the doping of Mn atoms can regulate the d-band center of the Ru active site in RuO2, lower antibonding Ru-O states, and enhance the intrinsic activity of RuO2 [23]. Hao et al. studied W and Er co-doping RuO2 lattice, and the results showed that the oxygen vacancies were increased, and the excessive oxidation of Ru was avoided [17]. Meanwhile, some work reported that the introduction of Sn atoms activates adjacent Ru sites and enables synthesized Sn/RuO2 nanoparticles to achieve high activity and durability [24].
Although great progress has been made regarding the OER catalytic activity of metal-doped RuO2, there is still room for improvement. Rutile RuO2 has always been considered a Pauli paramagnetic system, and the majority of studies treat it as a nonmagnetic system [25,26,27,28,29]. However, recent experimental measurements uncovered that its metallic state is linear antiferromagnetic [30,31]. In addition, Miao et al. recently demonstrated that the d-electron correlation of RuO2 with antiferromagnetic arrangement can tune the adsorption energy of oxygenated species, thereby affecting the OER catalytic performance [32]. Furthermore, the selection of active sites is also crucial to the evaluation of catalytic performance, while most of the previous studies have focused on the Ru atoms near doped atoms [4,15,23,33,34,35], which, to a certain extent, lack the exploration of active sites at other locations.
In this work, we comprehensively studied the OER catalytic performance of different active sites on Sn-doped antiferromagnetic RuO2 (Sn/RuO2)(110) and ( 1 ¯ 10) surfaces based on density functional theory (DFT). The stable geometric structure and atomic magnetic distribution of Sn/RuO2 bulk and surfaces are determined. Based on this, the adsorption energy and linear relationship of reaction intermediates *OH, *O, and *OOH at various active sites were calculated, and then the OER active volcano curve was plotted. It was found that the doping of Sn atoms effectively improved the OER activity of RuO2, and the OER performance of coordinatively unsaturated metal (Mcus) sites is generally higher than that of the bridge-bonded lattice oxygen (Obr) sites. Moreover, the site nearest to the doped Sn atom does not have the best OER activity. Finally, the effects of the d/p-band center, the average electrostatic potential of the ions and the charge transfer on the OER catalytic performance of Sn/RuO2 surface were further analyzed, which revealed the conditions for the optimal active site.

2. Results and Discussion

2.1. Stable Models of Sn/RuO2 Bulk and Surfaces

A reasonable Hubbard U value can accurately describe the electronic correlation in RuO2. As shown in Figure 1, RuO2 bulk is a nonmagnetic structure without considering U, while when the Ueff value is about 2 eV, the antiferromagnetic distribution RuO2 structure is more stable, agreeing with previous studies [30,32]. As Ueff increases to 4 eV, there is a band gap (0.47 eV) near the Fermi level, and RuO2 is no longer metallic, which is inconsistent with the actual properties of the material. Therefore, subsequent structure and activity studies are based on the Ueff value of 2 eV.
As shown in Figure 2a, a p(2 × 2 × 2) RuO2 supercell was modeled, in which, when 16 Ru atoms are randomly replaced by 2 Sn atoms, there will be 120 possible configurations. After screening for unequal symmetry, all the doping systems that have only six independent space groups, namely, Cmmm, Cm, P2/m, C2/m, Pmmm, and Fmmm (Figure S1), are found. So, it is necessary to calculate the six doping structures with different space groups and then select the most stable model as the system we focus on. Doped Sn atoms are located at the original Ru position, and their initial spin settings are the same as that of the original Ru atoms. As shown in Table S1, we consider the different spin configurations of each system. When two Sn atoms are in the nearest neighbor and their spin direction is opposite (Figure 2b), the Sn/RuO2 structure (Cm) with the most negative formation energy is thermodynamically stable.
Based on this, we explored the OER catalytic performance of Sn/RuO2 along the [110] direction. The schematic diagram of Figure 2c shows that different atomic distributions of Sn/RuO2 slabs in the [110] and [ 1 ¯ 10] directions will result in different exposed surfaces and active sites. Especially for the Sn/RuO2( 1 ¯ 10) surface, exposed surfaces with and without doped Sn atoms appear alternately. These surfaces are composed of the vertically oriented octahedral rows (singly coordinated O atoms at the surface) and the horizontally oriented octahedral rows (doubly coordinated O atoms at the surface). This work focuses on the OER catalytic activity of the Mcus and Obr sites on the Sn/RuO2 surfaces. Finally, the active sites and atomic distributions on the four unequal exposed surfaces of Sn/RuO2(110) and ( 1 ¯ 10) are shown in Figure 3a–d.

2.2. Reaction Mechanism and Free Energy Diagram

OER in acidic media usually follows a four-electron reaction path (Equations (1)–(4)), and H2O molecules deprotonate in turn to form *OH, *O, and *OOH intermediates, and then *OOH is further converted into product O2.
* + H 2 O   ( l )   * O H + H + + e   ( Δ G 1 = G * O H )
* O H * O + H + + e   ( Δ G 2 = G * O G * O H )
* O + H 2 O   ( l ) * O O H + H + + e   ( Δ G 3 = G * O O H G * O )
* O O H * + O 2 ( g ) + H + + e   ( Δ G 4 = 4.92 G * O O H )
where * represents the active site on the catalyst surface; Δ G 1 4 represents the reaction free energy of each step; and Δ G * O H , Δ G * O , and Δ G * O O H represent the adsorption free energy of *OH, *O, and *OOH intermediates, respectively, which can be obtained through Equations (5)–(7). The zero-point energy and entropy of the related gaseous molecules are displayed in Table S2.
G * O H = G * O H + 1 / 2 G H 2 G H 2 O E *
G * O = G * O H + G H 2 G H 2 O E *
G * O O H = G * O O H + 3 / 2 G H 2 2 G H 2 O E *
Following the above reaction steps, the adsorption of *OH, *O, and *OOH intermediates (Figure S2) was calculated. The OER reaction free energy diagram at equilibrium potential (1.23 V) and relaxed structures at the Mcus and Obr sites of Sn/RuO2 are shown in Figure 4, and the adsorption configuration of the remaining sites is similar. The Sabatier principle indicates that the activity of the catalyst is related to the adsorption intensity of the reaction intermediate [36,37]. Too strong an adsorption will hinder the desorption of the intermediate, resulting in the occupation of active sites; while too weak an adsorption will be unfavorable for site activation. Therefore, an ideal catalyst has a moderate binding strength with the intermediate. Generally, the elementary reaction step with the maximum reaction free energy is the rate-determining step (RDS), which significantly affects catalytic activity. The RDS of each site on the Sn/RuO2(110) and ( 1 ¯ 10) surfaces is mainly the process of *O formation and the conversion of *O into *OOH.
In Figure 4a, some sites bind too strongly to *OH, such as the 26-Obr site bridging Ru-Sn atoms (26-Obr_Ru-Sn), the 27-Obr_Ru-Ru site, the 12-Rucus site, and the 14-Rucus site, making the subsequent transformation of *O to *OOH difficult; whereas the adsorption of *O at the 11-Sncus site is so weak that there is a large RDS (*OH → *O). In contrast, the Δ G of *OH → *O at the 25-Rucus site is smaller and exhibits excellent OER catalytic activity on the Sn/RuO2(110) surface. Figure 4b shows the free energy diagrams of OER on the surface of Sn/RuO2( 1 ¯ 10). Similarly, a higher Δ G * O at the 31-Sncus site is unfavorable to OER progression. When lattice oxygen participates in the reaction, the adsorption of *OH and *O at the 33-Obr, 35-Obr, and 47-Obr sites is too strong, which, to some extent, hinders the formation of *OOH and its deprotonation process. However, the adsorption of oxygenated species at the 34-Rucus and 44-Rucus sites is moderate, displaying outstanding OER catalytic activity. The results are consistent with the previous study [24], where the Sncus sites exhibit the RDS at *OH → *O, while the Rucus sites show the RDS at *O → *OOH due to the strong *O binding strength.

2.3. Active Volcano Plot of OER Theory Overpotential

Theoretical overpotential ( η O E R ) is an important parameter in evaluating the OER activity of a catalyst. A lower η O E R value generally means better OER catalytic performance.
η O E R   = max { Δ G 1 ,   Δ G 2 ,   Δ G 3 ,   Δ G 4 } / e 1.23 V
From the definition of the η O E R , it can be seen that in order to achieve a lower η O E R , the Δ G i of each step should be close to 1.23 eV; that is, η O E R tends to 0 V, thereby balancing the binding strength of the reaction intermediates so that they are not too strong or too weak. Since OER activity is mainly determined by the Δ G * O O H , Δ G * O , and Δ G * O H , and they have a linear relationship with each other (Figure 5a), this simplifies the description of OER activity in the system.
As shown in Figure 5b, η O E R and Δ G * O Δ G * O H exhibit an approximate activity volcano plot, which is widely used for screening and developing efficient electrocatalysts. The Obr sites in the four exposed surfaces studied are concentrated on the left branch of the volcano curve, while the Mcus sites are mainly located on the right branch of the volcano plot, which can be attributed to the relatively weak binding of the Mcus site to *O than the Obr sites. Figure 5b displays that more unsaturated Rucus sites occupy the upper part of the volcano plot, where the 46-Rucus site (0.51 V) located at the top has the best OER catalytic activity. Then, the activity is followed by the 34- (0.58 V), 48- (0.80 V), 25- (0.82 V), and 14- (0.89 V) Rucus sites. The binding strength of *O and *OH at these sites is moderate, in line with the Sabatier principle.
In addition, we also calculated the OER catalytic performance of the pure RuO2(110) surface at the Rucus active sites before and after considering Hubbard U, and the corresponding reaction free energy diagram is shown in Figure S3. Before considering Ueff, the η O E R of RuO2(110) is 0.87 V, consistent with the previous results [38]. After considering Ueff, the η O E R of the Rucus site on the RuO2(110) surface is reduced by 0.14 eV, and its OER catalytic activity is enhanced. Compared with the η O E R of 46-Rucus site on the Sn/RuO2( 1 ¯ 10) surface, we can conclude that Sn doping activates the Ru-O bond and further improves the catalytic performance of OER.

2.4. Analysis of Electronic Properties on Sn/RuO2(110) and (110) Surfaces

The possible factors underlying the differences in the activity exhibited by various sites on the Sn/RuO2(110) and ( 1 ¯ 10) surfaces are further explored. The distance of Mcus or Obr active sites from the nearest doped Sn atom (dM/O-Sn) on the Sn/RuO2 surfaces is counted. For example, the distance between the 32-Rucus and 31-Sncus site is measured in Figure 3c, while the 34-Rucus site is counted as its distance from the nearest neighbor Sn atom instead of the 31-Sncus site. As shown in Figure 6a, Sn doping effectively enhanced the OER catalytic activity of RuO2, such as 46-Rucus, 32-Rucus, and 48-Rucus sites. However, in the case of either Mcus or Obr sites (Figure S4), it is not commonly assumed that the site closest to the doped atom is most active. When the value of dM/O-Sn at the active sites is approximately 3.6 Å, they exhibit outstanding OER catalytic performance. For transition metal oxide catalysts, the band center (ε) of the active atom is closely related to the binding strength of the adsorbed oxygenated species and the catalyst surface. Usually, a lower band center allows for the occupation of a high-energy anti-bonding state, resulting in a weak adsorption [39]. The active sites involved include Ru, Sn, and O atoms, so we explored the relationship between the Ru-d, Sn-p, and O-p band centers ( ε p / d ) and the η O E R .
As shown in Figure 6b, the ε p / d of the active site of the Sn/RuO2(110) and ( 1 ¯ 10) surfaces has a clear linear relationship with η O E R , and the closer the ε p / d is to the Fermi level, the better the OER catalytic activity at this site. However, it can also be observed from Figure 6b that the η O E R vs. ε p of the Sncus sites deviates from the trend line, which may be due to the intrinsic properties of the Sn atom located in the main group and the transition metal Ru atom located in the subgroup being quite different. To confirm this hypothesis, we calculated the ε p of the antimony (Sb) atom near the Sn atom, the ε d of the rhodium (Rh) atom around the Ru atom on the periodic table, and the OER catalytic activity of either Sb or Rh doped with RuO2(110), respectively (Figure S5). The data point of the 11-Sbcus site in Figure 6 is around the Sn site, which is also far away from the trend line; while Rh/RuO2 (purple data point) is located around the RuO2 system, which conforms to the trend line. Similar to Sn/RuO2, the 14-Rucus site (dRu-Sb/Rh = 3.6 Å) on the Sb (Rh)/RuO2(110) surface also displays optimal OER performance (Figure 6c). Slightly different from the Sn and Sb atoms, the Rh site exhibits OER activity comparable to that of the 14-Rucus site, mainly due to the similar valence electron distribution of Rh and Ru atoms. The electron density of states (DOS) for Sb/RuO2 and Rh/RuO2 is further compared (Figure 6d). In the Sb/RuO2(110) surface, the electronic states near the Fermi level are predominantly contributed by the 12-Rucus and 14-Rucus sites. As for the Rh/RuO2, the 11-Rh site exhibits a higher electron occupancy state near the Fermi level, which accounts for the superior activity of the 11-Rhcus site compared to the 11-Sbcus site.
Moreover, we analyzed the contributions of Bader charge (red data points) and average electrostatic potential at the core (blue data points) of metal atoms at each active site to their catalytic activity. Figure 7a–d shows the Bader charge distribution of each atom on the Sn/RuO2(110) and ( 1 ¯ 10) surfaces, where the Sn and Ru atoms donate electrons to the O atoms. The doping of Sn atoms increases the electron transfer of neighboring Ru atoms compared to the distant Ru sites, thereby affecting the adsorption intensity and catalytic activity of the Rucus sites. The quantitative analysis is displayed in Figure 7e,f. It is noted that the Ueff correction significantly affects the electronic properties of RuO2 system. In Figure 7e, the RuO2(110) surface with Ueff loses more electrons and exhibits a lower average electrostatic potential than the RuO2(110) surface without Ueff. Usually Ru-based catalysts are easily over-oxidized into unstable high valence state Ru in acidic electrolytes, resulting in the gradual dissolution of the materials [17]. However, the less charge is transferred from 46-Rucus and 34-Rucus sites than the original undoped RuO2. This indicates that Sn doping not only enhances the OER activity but also reduces the oxidation state of partial Ru sites and increases their antioxidant properties. Du et al. also indicated that electron enrichment on Ru ions can decreases oxidation and maintain the stability of RuO2 in the hybrid catalyst [40]. The 12-Rucus and 32-Rucus sites closest to the Sn atom transfer more electrons than the other Rucus sites and possess a high overpotential. Overall, lower OER activity at the Rucus site correlates with more negative electrostatic potential and higher charge transfer.
As for the Obr sites, given that lattice O binds with OH to release O2, the reaction intermediates adsorb at the Obr site and the oxygen vacancy (Ov). Thus, the electrostatic potential and Bader charge of two metal atoms near the Ov are taken into account, and the connection between their geometric mean and η O E R is established (Figure 7f). Similar to the trend of the Rucus sites, the OER performance at the Obr sites declines with increasingly negative electrostatic potentials or enhanced electron transfer on adjacent metal atoms. However, an opposite trend is observed at the doped Sncus and the Obr_Ru-Sn sites. In Figure S6, more electrons are lost at the sites closer to the Sn atom, and their atomic electrostatic potential is significantly lower than that of the Rucus and the Obr_Ru-Ru sites. As the atomic electrostatic potential and Bader charge gradually decrease, it seems to be more favorable for the OER reaction at the site containing the Sn atom.
To better clarify the role of Sn doping, in the case of Sn/RuO2( 1 ¯ 10) surface, the Crystal Orbital Hamilton Population (COHP) [41,42] is calculated to analyze the bonding interactions between active sites and *O intermediate, including undoped RuO2 with Ueff, 31-Sncus, 32-Rucus, and 34-Rucus sites. The ICOHP value obtained from the COHP integration indicates the bond strength between atomic pairs; i.e., the more negative the ICOHP, the stronger binding strength. As shown in Figure 8, the order of the ICOHP at various sites is RuO2 < 34-Sncus < 32-Rucus < 31-Rucus. The adsorption of *O on the undoped RuO2 is too strong, hindering the conversion of *O to *OOH. The weak adsorption of *O at the 31-Sncus site results in a RDS of *OOH → *O. However, the ICOHP at the 34-Rucus site is in the middle, and the adsorption strength of *O is moderate. The introduced Sn atoms adjust the original adsorption strength of the oxygenated species, thereby enhancing the electrocatalytic performance of RuO2. In addition, the differential charge density of O2 on the RuO2(110) and Sn/RuO2(110) surfaces also proves the role of Sn (Figure S7). The Ru atoms donate electrons to O2 for the formation of Ru-O bonds. For undoped RuO2, the electrons transferred by the Ru atoms connected to O atom are 1.74 e, whereas for the Sn/RuO2 surface, this value decreases to 1.69 e. It is clear that Sn doping modulates the electron distribution of neighboring Ru atoms, weakening the binding of Rucus sites to O2, favoring O2 desorption, and thereby accelerating the OER process.

3. Computational Details

All spin polarization DFT calculations in this work were implemented by the Vienna Ab-initio Simulation Package (VASP) [43,44]. The projector augmented wave (PAW) pseudopotential [45] was used to model the interaction between the ions and electrons, and a 520 eV cutoff energy of the plane-wave was set. The Perdew–Burke–Ernzerhof (PBE) functional in the GGA approximation framework was adopted to describe the exchange correlation of electrons [46]. The Hubbard U method was used to deal with strongly correlated d-electronic systems [47], and an effective Ueff value of 2 eV was used for the 3d electrons in Ru atoms [30,32]. Structural relaxation is performed until the force on each atom is below 0.02 eV/Å, and the energy is converged within 10−5 eV. A p(2 × 2 × 2) rutile Ru16O32 and a Sn2Ru14O32 bulk structure with two Ru atoms replaced by Sn (referred to as Sn/RuO2) were studied, in which a 3 × 3 × 5 Gamma-centered k-point grid was applied. A stoichiometric RuO2(110) (including 64 Ru atoms and 128 O atoms), Sn/RuO2(110), and Sn/RuO2( 1 ¯ 10) slab models (composed of 56 Ru atoms, 8 Sn atoms, and 128 O atoms) were constructed to explore the effects of Sn doping. A vacuum layer with 15 Å thickness was used to avoid the interaction between periodic slabs in the direction perpendicular to the surface, and a 2 × 2 × 1 k-point grid was used for these surfaces. During the calculation process, the bottom four atomic layers of slab were fixed while the other atomic layers were sufficiently relaxed. The initial magnetic moment distribution of metal atoms in RuO2(110), Sn/RuO2(110), and Sn/RuO2( 1 ¯ 10) slabs was consistent with the antiferromagnetic arrangement in their bulk structures before surface cleaving.
A negative formation energy (Ef) usually means that the system is thermodynamically stable. Thus, the E f b u l k was calculated using Equation (9) to determine the stability of Sn/RuO2 bulk structures.
E f b u l k = E S n / R u O 2 b u l k + 2 E R u b u l k ( E R u O 2 b u l k + 2 E S n b u l k )
where E R u O 2 b u l k and E S n / R u O 2 b u l k are the total energy of the systems before (Ru16O32) and after (Sn2Ru14O32) Sn doping, respectively. E R u b u l k and E S n b u l k denote the energy of the individual Ru and Sn atom in their metal bulk crystal. In addition, the Gibbs free energy (G) of the system is defined as follows:
G = E D F T + E Z P E T S
in which E D F T is the total energy of the systems calculated by the VASP; and T is the room temperature 273.15 K. E Z P E and S are the zero-point energy and entropy, which can be obtained from the vibration frequency of the adsorbed species.

4. Conclusions

In summary, we systematically investigated the OER electrocatalytic activity of Mcus and Obr sites on different exposed surfaces of Sn/RuO2(110) and ( 1 ¯ 10) based on the DFT+U method. The results demonstrate that the magnetic properties of the RuO2 structure corrected by Hubbard U change from the original paramagnetic to antiferromagnetic, and the OER overpotential of the RuO2(110) surface is reduced by 0.14 V. According to the linear relationship between the reaction intermediates *OOH, *O, and *OH, the OER active volcano plot of the Sn/RuO2(110) and ( 1 ¯ 10) surfaces are constructed. The activity of the Mcus sites is generally higher than that of the Obr sites, where the 46-Rucus site possesses the optimal OER catalytic performance, followed by the 34-Rucus < 48-Rucus < 25-Rucus < 14-Rucus sites. Moreover, contrary to conventional knowledge, the active sites with a distance of about 3.6 Å from the doped Sn atom exhibit a lower OER overpotential rather than the nearest neighbor sites. By studying the electronic properties of various active sites on the Sn/RuO2 surface, the intrinsic factors affecting activity are further uncovered, such as the center of the d/p band, the electrostatic potential of the atomic core, and the number of electrons transferred. The evaluation of catalytic performance is highly dependent on the correct selection of the active site. This study plays an important role in the simulation and design of high-performance electrocatalysts for energy conversion and storage.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal15080770/s1: Figure S1: Bulk structures of two Sn atoms doped at different sites in the RuO2 system: (a) Cmmm, (b) Cm, (c) P2/m, (d) C2/m, (e) Pmmm, and (f) Fmmm space group; Figure S2: Relaxation structures of *OH, *O, and *OOH at the (a-d) Mcus and (e–h) Obr sites: (a) sub, (b) *OH, (c) *O, (d) *OOH; (e) sub-Ov, (f) *ObrH, (g) *Obr, (h) *ObrOH; Figure S3: Reaction free energy diagram of RuO2(110) surface without and with Hubbard U (Ueff = 2 eV); Figure S4: OER activity distribution of (a) Mcus sites and (b) Obr sites at different distances from the doped Sn atom; Figure S5: Free energy diagrams of (a) Sb/RuO2(110) and (b) Rh/RuO2(110) at electrode potential U = 1.23 V; Figure S6: Relationship between electrostatic potential, d/p band center, Bader charge, and OER catalytic activity: (a) Sncus and Obr_Ru-Sn sites in Sn/RuO2 system; (b) Mcus and Obr sites in Sb (Rh)/RuO2(110) surface; Figure S7: Top and side view of the charge density difference of O2 adsorbed on the RuO2(110) surface without (a) and with (b) Sn doping; Table S1: Total energy difference and formation energy ( E f b u l k , eV) of Sn/RuO2 bulk structure under different magnetic moments (Mag, μB); Table S2: Zero-point energy (EZPE) and entropy (S) correction of the reaction intermediates in free energy calculation.

Author Contributions

Conceptualization, Z.H. and C.Z.; methodology, Z.H. and C.Z.; software, C.Z.; validation, C.Z. and Z.H.; formal analysis, C.Z., Q.G. and Z.H.; investigation, C.Z.; resources, Z.H.; writing—original draft preparation, C.Z. and Z.H.; writing—review and editing, C.Z. and Z.H.; visualization, C.Z.; supervision, Z.H.; project administration, Z.H.; funding acquisition, Z.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 21773124, 21933006), the Open Fund of Key Laboratory of Advanced Electrode Materials for Novel Solar Cells for Petroleum and Chemical Industry of China (No. 2024A089), the Fundamental Research Funds for the Central Universities Nankai University (No. 63243091, 63213042), and the Supercomputing Center of Nankai University (NKSC).

Data Availability Statement

The original contributions presented in this study are included in the article and Supplementary Materials. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chu, S.; Majumdar, A. Opportunities and challenges for a sustainable energy future. Nature 2012, 488, 294–303. [Google Scholar] [CrossRef]
  2. Lin, Y.; Dong, Y.; Wang, X.; Chen, L. Electrocatalysts for the Oxygen Evolution Reaction in Acidic Media. Adv. Mater. 2023, 35, 2210565. [Google Scholar] [CrossRef]
  3. Hu, S.; Ge, S.; Liu, H.; Kang, X.; Yu, Q.; Liu, B. Low-Dimensional Electrocatalysts for Acidic Oxygen Evolution: Intrinsic Activity, High Current Density Operation, and Long-Term Stability. Adv. Funct. Mater. 2022, 32, 2201726. [Google Scholar] [CrossRef]
  4. Zhang, Y.; Dong, J.; Sun, T.; Zhang, X.; Chen, J.; Xu, L. Mo-Doped Mesoporous RuO2 Spheres as High-Performance Acidic Oxygen Evolution Reaction Electrocatalyst. Small 2024, 20, 2305889. [Google Scholar] [CrossRef]
  5. Li, L.; Wang, P.; Shao, Q.; Huang, X. Recent Progress in Advanced Electrocatalyst Design for Acidic Oxygen Evolution Reaction. Adv. Mater. 2021, 33, 2004243. [Google Scholar] [CrossRef]
  6. Zhai, P.; Xia, M.; Wu, Y.; Zhang, G.; Gao, J.; Zhang, B.; Cao, S.; Zhang, Y.; Li, Z.; Fan, Z.; et al. Engineering single-atomic ruthenium catalytic sites on defective nickel-iron layered double hydroxide for overall water splitting. Nat. Commun. 2021, 12, 4587. [Google Scholar] [CrossRef]
  7. Wang, Y.; Yang, R.; Ding, Y.; Zhang, B.; Li, H.; Bai, B.; Li, M.; Cui, Y.; Xiao, J.; Wu, Z.-S. Unraveling oxygen vacancy site mechanism of Rh-doped RuO2 catalyst for long-lasting acidic water oxidation. Nat. Commun. 2023, 14, 1412. [Google Scholar] [CrossRef]
  8. Dickens, C.F.; Nørskov, J.K. A Theoretical Investigation into the Role of Surface Defects for Oxygen Evolution on RuO2. J. Phys. Chem. C 2017, 121, 18516–18524. [Google Scholar] [CrossRef]
  9. Hu, C.; Hu, Y.; Zhang, B.; Zhang, H.; Bao, X.; Zhang, J.; Yuan, P. Advanced Catalyst Design Strategies and In-Situ Characterization Techniques for Enhancing Electrocatalytic Activity and Stability of Oxygen Evolution Reaction. Electrochem. Energy Rev. 2024, 7, 19. [Google Scholar] [CrossRef]
  10. Shi, Q.; Zhu, C.; Du, D.; Lin, Y. Robust noble metal-based electrocatalysts for oxygen evolution reaction. Chem. Soc. Rev. 2019, 48, 3181–3192. [Google Scholar] [CrossRef]
  11. Man, I.C.; Su, H.-Y.; Calle-Vallejo, F.; Hansen, H.A.; Martínez, J.I.; Inoglu, N.G.; Kitchin, J.; Jaramillo, T.F.; Nørskov, J.K.; Rossmeisl, J. Universality in Oxygen Evolution Electrocatalysis on Oxide Surfaces. ChemCatChem 2011, 3, 1159–1165. [Google Scholar] [CrossRef]
  12. Gao, H.; Xu, J.; Zhang, X.; Zhou, M. Benchmarking the Intrinsic Activity of Transition Metal Oxides for the Oxygen Evolution Reaction with Advanced Nanoelectrodes. Angew. Chem. Int. Ed. 2024, 63, e202404663. [Google Scholar] [CrossRef]
  13. Wang, A.; Li, J.; Zhang, T. Heterogeneous single-atom catalysis. Nat. Rev. Chem. 2018, 2, 65–81. [Google Scholar] [CrossRef]
  14. Qiao, B.; Wang, A.; Yang, X.; Allard, L.F.; Jiang, Z.; Cui, Y.; Liu, J.; Li, J.; Zhang, T. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nature Chem. 2011, 3, 634–641. [Google Scholar] [CrossRef]
  15. Tian, Y.; Wang, S.; Velasco, E.; Yang, Y.; Cao, L.; Zhang, L.; Li, X.; Lin, Y.; Zhang, Q.; Chen, L. A Co-Doped Nanorod-like RuO2 Electrocatalyst with Abundant Oxygen Vacancies for Acidic Water Oxidation. iScience 2020, 23, 100756. [Google Scholar] [CrossRef]
  16. Sardar, K.; Petrucco, E.; Hiley, C.I.; Sharman, J.D.B.; Wells, P.P.; Russell, A.E.; Kashtiban, R.J.; Sloan, J.; Walton, R.I. Water-Splitting Electrocatalysis in Acid Conditions Using Ruthenate-Iridate Pyrochlores. Angew. Chem. Int. Ed. 2014, 53, 10960–10964. [Google Scholar] [CrossRef]
  17. Hao, S.; Liu, M.; Pan, J.; Liu, X.; Tan, X.; Xu, N.; He, Y.; Lei, L.; Zhang, X. Dopants fixation of Ruthenium for boosting acidic oxygen evolution stability and activity. Nat. Commun. 2020, 11, 5368. [Google Scholar] [CrossRef]
  18. Nong, H.N.; Oh, H.-S.; Reier, T.; Willinger, E.; Willinger, M.-G.; Petkov, V.; Teschner, D.; Strasser, P. Oxide-Supported IrNiOx Core–Shell Particles as Efficient, Cost-Effective, and Stable Catalysts for Electrochemical Water Splitting. Angew. Chem. Int. Ed. 2015, 54, 2975–2979. [Google Scholar] [CrossRef]
  19. Qin, Y.; Yu, T.; Deng, S.; Zhou, X.-Y.; Lin, D.; Zhang, Q.; Jin, Z.; Zhang, D.; He, Y.-B.; Qiu, H.-J.; et al. RuO2 electronic structure and lattice strain dual engineering for enhanced acidic oxygen evolution reaction performance. Nat. Commun. 2022, 13, 3784. [Google Scholar] [CrossRef]
  20. Wang, C.; Jin, L.; Shang, H.; Xu, H.; Shiraishi, Y.; Du, Y. Advances in engineering RuO2 electrocatalysts towards oxygen evolution reaction. Chin. Chem. Lett. 2021, 32, 2108–2116. [Google Scholar] [CrossRef]
  21. Zhou, Y.-N.; Yu, N.; Lv, Q.-X.; Liu, B.; Dong, B.; Chai, Y.-M. Surface evolution of Zn doped-RuO2 under different etching methods towards acidic oxygen evolution. J. Mater. Chem. A 2022, 10, 16193–16203. [Google Scholar] [CrossRef]
  22. Ede, S.R.; Bijoy, T.K.; Sankar, S.S.; Murugan, P.; Kundu, S. Rational Design of Highly Efficient Perovskite Hydroxide for Electrocatalytic Water Oxidation. Inorg. Chem. 2020, 59, 4816–4824. [Google Scholar] [CrossRef]
  23. Chen, S.; Huang, H.; Jiang, P.; Yang, K.; Diao, J.; Gong, S.; Liu, S.; Huang, M.; Wang, H.; Chen, Q. Mn-Doped RuO2 Nanocrystals as Highly Active Electrocatalysts for Enhanced Oxygen Evolution in Acidic Media. ACS Catal. 2020, 10, 1152–1160. [Google Scholar] [CrossRef]
  24. Qiu, L.; Zheng, G.; He, Y.; Lei, L.; Zhang, X. Ultra-small Sn-RuO2 nanoparticles supported on N-doped carbon polyhedra for highly active and durable oxygen evolution reaction in acidic media. Chem. Eng. J. 2021, 409, 128155. [Google Scholar] [CrossRef]
  25. Ryden, W.D.; Lawson, A.W. Magnetic Susceptibility of IrO2 and RuO2. J. Chem. Phys. 1970, 52, 6058–6061. [Google Scholar] [CrossRef]
  26. Lin, Y.; Tian, Z.; Zhang, L.; Ma, J.; Jiang, Z.; Deibert, B.J.; Ge, R.; Chen, L. Chromium-ruthenium oxide solid solution electrocatalyst for highly efficient oxygen evolution reaction in acidic media. Nat. Commun. 2019, 10, 162. [Google Scholar] [CrossRef]
  27. Riga, J.; Tenret-Noël, C.; Pireaux, J.J.; Caudano, R.; Verbist, J.J.; Gobillon, Y. Electronic Structure of Rutile Oxides TiO2, RuO2 and IrO2 Studied by X-ray Photoelectron Spectroscopy. Phys. Scr. 1977, 16, 351. [Google Scholar] [CrossRef]
  28. Saha, S.; Kishor, K.; Pala, R.G.S. Increasing electrochemical chlorine selectivity over oxygen selectivity through the optimal weakening of oxygen bonds in transition metal-doped RuO2. Catal. Sci. Technol. 2024, 14, 4566–4574. [Google Scholar] [CrossRef]
  29. Liu, S.; Chang, Y.; He, N.; Zhu, S.; Wang, L.; Liu, X. Competition between Lattice Oxygen and Adsorbate Evolving Mechanisms in Rutile Ru-Based Oxide for the Oxygen Evolution Reaction. ACS Appl. Mater. Interfaces 2023, 15, 20563–20570. [Google Scholar] [CrossRef]
  30. Berlijn, T.; Snijders, P.C.; Delaire, O.; Zhou, H.D.; Maier, T.A.; Cao, H.B.; Chi, S.X.; Matsuda, M.; Wang, Y.; Koehler, M.R.; et al. Itinerant Antiferromagnetism in RuO2. Phys. Rev. Lett. 2017, 118, 077201. [Google Scholar] [CrossRef]
  31. Zhu, Z.H.; Strempfer, J.; Rao, R.R.; Occhialini, C.A.; Pelliciari, J.; Choi, Y.; Kawaguchi, T.; You, H.; Mitchell, J.F.; Shao-Horn, Y.; et al. Anomalous Antiferromagnetism in Metallic RuO2 Determined by Resonant X-ray Scattering. Phys. Rev. Lett. 2019, 122, 017202. [Google Scholar] [CrossRef]
  32. Miao, X.; Zhang, J.; Hu, Z.; Zhou, S. Modulating Electronic Correlations in Ruthenium Oxides for Highly Efficient Oxygen Evolution Reaction. Precis. Chem. 2025, 3, 72–81. [Google Scholar] [CrossRef]
  33. Feng, T.; Yu, J.; Yue, D.; Song, H.; Tao, S.; Waterhouse, G.I.N.; Lu, S.; Yang, B. Defect-rich ruthenium dioxide electrocatalyst enabled by electronic reservoir effect of carbonized polymer dot for remarkable pH-universal oxygen evolution. Appl. Catal. B Environ. 2023, 328, 122546. [Google Scholar] [CrossRef]
  34. Zhang, L.; Jang, H.; Liu, H.; Kim, M.G.; Yang, D.; Liu, S.; Liu, X.; Cho, J. Sodium-Decorated Amorphous/Crystalline RuO2 with Rich Oxygen Vacancies: A Robust pH-Universal Oxygen Evolution Electrocatalyst. Angew. Chem. Int. Ed. 2021, 60, 18821–18829. [Google Scholar] [CrossRef]
  35. Wang, K.; Wang, Y.; Yang, B.; Li, Z.; Qin, X.; Zhang, Q.; Lei, L.; Qiu, M.; Wu, G.; Hou, Y. Highly active ruthenium sites stabilized by modulating electron-feeding for sustainable acidic oxygen-evolution electrocatalysis. Energy Environ. Sci. 2022, 15, 2356–2365. [Google Scholar] [CrossRef]
  36. Hu, S.; Li, W.-X. Sabatier principle of metal-support interaction for design of ultrastable metal nanocatalysts. Science 2021, 374, 1360–1365. [Google Scholar] [CrossRef]
  37. Sabatier, P. Hydrogénations et déshydrogénations par catalyse. Ber. Dtsch. Chem. Ges. 1911, 44, 1984–2001. [Google Scholar] [CrossRef]
  38. Miao, X.; Zhang, L.; Wu, L.; Hu, Z.; Shi, L.; Zhou, S. Quadruple perovskite ruthenate as a highly efficient catalyst for acidic water oxidation. Nat. Commun. 2019, 10, 3809. [Google Scholar] [CrossRef]
  39. Hammer, B.; Norskov, J.K. Why gold is the noblest of all the metals. Nature 1995, 376, 238–240. [Google Scholar] [CrossRef]
  40. Du, K.; Zhang, L.; Shan, J.; Guo, J.; Mao, J.; Yang, C.-C.; Wang, C.-H.; Hu, Z.; Ling, T. Interface engineering breaks both stability and activity limits of RuO2 for sustainable water oxidation. Nat. Commun. 2022, 13, 5448. [Google Scholar] [CrossRef]
  41. Dronskowski, R.; Bloechl, P.E. Crystal orbital Hamilton populations (COHP): Energy-resolved visualization of chemical bonding in solids based on density-functional calculations. J. Phys. Chem. 1993, 97, 8617–8624. [Google Scholar] [CrossRef]
  42. Deringer, V.L.; Tchougréeff, A.L.; Dronskowski, R. Crystal Orbital Hamilton Population (COHP) Analysis As Projected from Plane-Wave Basis Sets. J. Phys. Chem. A 2011, 115, 5461–5466. [Google Scholar] [CrossRef]
  43. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef] [PubMed]
  44. 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, 15–50. [Google Scholar] [CrossRef]
  45. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed]
  46. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  47. Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, A.P. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study. Phys. Rev. B 1998, 57, 1505–1509. [Google Scholar] [CrossRef]
Figure 1. Properties of RuO2 bulk at different Hubbard U values: (ac) density of states in antiferromagnetic (AFM) configurations; (d) calculated total energy (Etot), magnetic moment (MRu), and band gap (Egap) of the AFM and nonmagnetic (NM) systems.
Figure 1. Properties of RuO2 bulk at different Hubbard U values: (ac) density of states in antiferromagnetic (AFM) configurations; (d) calculated total energy (Etot), magnetic moment (MRu), and band gap (Egap) of the AFM and nonmagnetic (NM) systems.
Catalysts 15 00770 g001
Figure 2. Schematic diagram of side and top view of (a) RuO2, (b) Sn/RuO2, and (c) surface atomic arrangement of RuO2 along the (110) and ( 1 ¯ 10) direction. The gray, yellow, purple, and pink balls represent spin-up Ru, spin-down Ru, Sn, and O atoms, respectively.
Figure 2. Schematic diagram of side and top view of (a) RuO2, (b) Sn/RuO2, and (c) surface atomic arrangement of RuO2 along the (110) and ( 1 ¯ 10) direction. The gray, yellow, purple, and pink balls represent spin-up Ru, spin-down Ru, Sn, and O atoms, respectively.
Catalysts 15 00770 g002
Figure 3. Top and side views of (a,b) Sn/RuO2(110) and (c,d) Sn/RuO2( 1 ¯ 10) surfaces. The numbers in the figure represent the different active sites. The gray, yellow, purple, and pink balls represent spin-up Ru, spin-down Ru, Sn, and O atoms, respectively.
Figure 3. Top and side views of (a,b) Sn/RuO2(110) and (c,d) Sn/RuO2( 1 ¯ 10) surfaces. The numbers in the figure represent the different active sites. The gray, yellow, purple, and pink balls represent spin-up Ru, spin-down Ru, Sn, and O atoms, respectively.
Catalysts 15 00770 g003
Figure 4. Free energy diagrams of (a) Sn/RuO2(110) and (b) Sn/RuO2( 1 ¯ 10) at electrode potential U = 1.23 V. Relaxed adsorption structures of oxygenated species at (c) Mcus and (d) Obr sites. The gray, yellow, purple, pink, and white balls represent spin-up Ru, spin-down Ru, Sn, O, and H atoms, respectively.
Figure 4. Free energy diagrams of (a) Sn/RuO2(110) and (b) Sn/RuO2( 1 ¯ 10) at electrode potential U = 1.23 V. Relaxed adsorption structures of oxygenated species at (c) Mcus and (d) Obr sites. The gray, yellow, purple, pink, and white balls represent spin-up Ru, spin-down Ru, Sn, O, and H atoms, respectively.
Catalysts 15 00770 g004
Figure 5. (a) Linear relationship between the adsorption free energies of *O, *OOH, and *OH. (b) OER overpotential as a function of G*O − ΔG*OH for Sn/RuO2(110) and ( 1 ¯ 10) surfaces.
Figure 5. (a) Linear relationship between the adsorption free energies of *O, *OOH, and *OH. (b) OER overpotential as a function of G*O − ΔG*OH for Sn/RuO2(110) and ( 1 ¯ 10) surfaces.
Catalysts 15 00770 g005
Figure 6. OER activity distribution and electronic structure: (a) Sn/RuO2(110) and Sn/RuO2( 1 ¯ 10) surfaces; (b) relationship between the OER overpotential and the p/d-band center of the Sn/RuO2 system; (c) OER activity at various sites on the Rh (Sb)/RuO2(110) surface; (d) density of states for the Sb (Rh)/RuO2(110) surface.
Figure 6. OER activity distribution and electronic structure: (a) Sn/RuO2(110) and Sn/RuO2( 1 ¯ 10) surfaces; (b) relationship between the OER overpotential and the p/d-band center of the Sn/RuO2 system; (c) OER activity at various sites on the Rh (Sb)/RuO2(110) surface; (d) density of states for the Sb (Rh)/RuO2(110) surface.
Catalysts 15 00770 g006
Figure 7. Calculated Bader charge distribution and electrostatic potential: (a,b) Sn/RuO2(110); (c,d) Sn/RuO2( 1 ¯ 10); (e) Rucus sites; (f) Obr sites.
Figure 7. Calculated Bader charge distribution and electrostatic potential: (a,b) Sn/RuO2(110); (c,d) Sn/RuO2( 1 ¯ 10); (e) Rucus sites; (f) Obr sites.
Catalysts 15 00770 g007
Figure 8. The COHP of (a) Rucus site in RuO2 structure, (b) 31-Sncus site, (c) 32-Rucus site, and (d) 34-Rucus sites in Sn/RuO2( 1 ¯ 10) structure. Gray and blue indicate spin-up and spin-down, respectively.
Figure 8. The COHP of (a) Rucus site in RuO2 structure, (b) 31-Sncus site, (c) 32-Rucus site, and (d) 34-Rucus sites in Sn/RuO2( 1 ¯ 10) structure. Gray and blue indicate spin-up and spin-down, respectively.
Catalysts 15 00770 g008
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

Zheng, C.; Gao, Q.; Hu, Z. First-Principles Study of Sn-Doped RuO2 as Efficient Electrocatalysts for Enhanced Oxygen Evolution. Catalysts 2025, 15, 770. https://doi.org/10.3390/catal15080770

AMA Style

Zheng C, Gao Q, Hu Z. First-Principles Study of Sn-Doped RuO2 as Efficient Electrocatalysts for Enhanced Oxygen Evolution. Catalysts. 2025; 15(8):770. https://doi.org/10.3390/catal15080770

Chicago/Turabian Style

Zheng, Caiyan, Qian Gao, and Zhenpeng Hu. 2025. "First-Principles Study of Sn-Doped RuO2 as Efficient Electrocatalysts for Enhanced Oxygen Evolution" Catalysts 15, no. 8: 770. https://doi.org/10.3390/catal15080770

APA Style

Zheng, C., Gao, Q., & Hu, Z. (2025). First-Principles Study of Sn-Doped RuO2 as Efficient Electrocatalysts for Enhanced Oxygen Evolution. Catalysts, 15(8), 770. https://doi.org/10.3390/catal15080770

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