Badr
Elkamash
and
Franziska
Hess
*
Institute for Chemistry, Technical University Berlin, D-10623 Berlin, Germany. E-mail: f.hess@tu-berlin.de
First published on 26th March 2025
To advance sustainable green hydrogen production through water electrolysis, where ruthenium dioxide (RuO2) is a promising catalyst for the oxygen evolution reaction (OER), we investigate the stability of RuO2, focusing on corrosion resistance – a key challenge limiting its practical application. Using density functional theory (DFT), we analyze the thermodynamic stability and reaction pathways across various RuO2 surface orientations, with a primary focus on the (110) surface. Specifically, we assess the impact of doping with Ta, W, Re, Ir, Ti, and Pt on the thermodynamic stability of the RuO2(110) surface against dissolution of Ru surface atoms in the form of RuO4. Our findings reveal that dopants Ir, Ti, and Pt in low oxidation states significantly enhance the resistance of the RuO2(110) surface against corrosion, while Ta, W, and Re in high oxidation states destabilize the surface, promoting degradation. We also identify specific dopant sites, such as those next to or directly underneath the dissolving Ru atom, that contribute significantly to surface stability, providing a roadmap for optimizing RuO2 catalysts. Additionally, we extend the investigation to reaction pathways towards the dissolution of the Ru atom by incorporating the effects of dopants, revealing that dopants not only alter the thermodynamic stability but also the reaction mechanism due to their different termination preferences. We establish a strong linear correlation between the Gibbs free energy of RuO4 formation (ΔGtot) and the free energy of the highest intermediate (ΔGmax), proposing ΔGtot as a reliable descriptor for predicting the thermodynamic stability of doped RuO2 surfaces against Ru dissolution. This allows for efficient computational screening of surface modifications, including dopant selection and surface orientation tuning, without requiring detailed knowledge of the entire stepwise mechanism toward the formation and removal of RuO4. This insight enables efficient computational screening of dopants and surface modifications, providing a framework for optimizing RuO2 catalysts to improve durability in electrochemical applications.
Addressing the challenges associated with the OER is paramount for advancing the commercial viability and widespread adoption of water electrolysis technology. Ruthenium dioxide (RuO2) has emerged as a promising catalyst for the OER, owing to its high activity and favorable electrochemical properties.4 Despite its potential, the practical application of RuO2 in water electrolysis is hindered by issues related to its stability under operating conditions.5 The dissolution of Ru from the RuO2 structure leads to the formation of soluble RuO4 species, compromising the long-term stability and performance of RuO2-based catalysts.6–8 This instability is particularly pronounced at surface defect sites and undercoordinated Ru atoms, where dissolution is more likely to occur. Understanding the atomic-scale mechanisms of RuO2 degradation is therefore essential for developing more stable catalyst formulations.
Enhancing the stability of RuO2 requires a fundamental understanding of the mechanisms governing its degradation and the development of effective strategies to mitigate corrosion. One promising approach is heteroatom doping, where the incorporation of stable elements such as Ir, Ti, and Pt can modify the electronic structure of RuO2, suppress Ru dissolution, and improve overall durability. For instance, a Ru0.5Ir0.5 alloy achieves four times higher stability compared to other Ru–Ir OER materials by adjusting surface composition through segregation.9 In Ir–Ru mixed oxides, Ru dissolves more rapidly than Ir, leading to activity loss, while the corrosion resistance of Ir contributes to overall stability.10 Titanium substitution in RuO2 at low concentrations (12.5% and 20% Ti) enhances catalyst stability and reduces Ru dissolution.11 TiRuO2 solid solutions improve selectivity towards cathodic electrochemical reactions (CER), such as the hydrogen evolution reaction (HER), rather than enhance OER activity.
Another approach involves oxygen vacancy engineering, where dopants such as W and Er increase the formation energy of oxygen vacancies, effectively improving the structural integrity of RuO2 and reducing dissolution in acidic PEM electrolyzers.12 Furthermore, surface modifications, such as atomic layer deposition (ALD) of protective coatings and the formation of mixed-metal oxide layers, have been shown to enhance RuO2 stability, with submonolayer IrOx on RuO2 significantly suppressing Ru dissolution and improving durability in PEM electrolyzers.13 Meanwhile, Ru nanoparticles exhibit severe corrosion and instability during OER, whereas Ir nanoparticles show improved durability, indicating their potential as effective nanoscale OER catalysts.14 Dispersing RuO2 over defective TiO2 enhances acidic OER performance by lowering the *OOH formation barrier and improving stability.15
In this study, leveraging Density Functional Theory (DFT) calculations in the computational hydrogen electrode model,16,17 we investigate the thermodynamic stability and reaction pathways of RuO2 catalysts in water electrolysis. We examine surface defects across various orientations, including (110), (100), c(2 × 2)-reconstructed, and step edge configurations, to gain insights into the influence of surface morphology on RuO2 stability against corrosion. Additionally, we explore the effects of support materials like M = (Ta, W, Re, Ir, Ti, and Pt), selected based on their stability in oxidative environments, electrochemical properties, and structural compatibility with RuO2 due to their ability to form MO2 rutile structures, thereby potentially forming stable solid solutions and interfaces with RuO2. These elements are known to influence corrosion resistance, making them ideal candidates for this study. Their effects, either as a substrate or sublayer, on RuO2 surface composition and stability are examined. Finally, we investigate how doping with a selection of these elements influences the thermodynamic stability of the RuO2(110) surface against corrosion.
In order to study the connection between thermodynamics and kinetics of corrosion, represented by the overall Gibbs free energy of RuO4 formation (ΔGtot) and the free energy of the highest intermediate (ΔGmax), we delve into the reaction pathways of RuO4 formation at the bridge site of the (110) surface. Here, we focus on how different dopants alter the corrosion mechanisms and overall stability of RuO2 surfaces. By detailing these reaction pathways, we aim to identify opportunities for optimizing RuO2 catalysts to enhance their performance in water electrolysis. Finally, we demonstrate that ΔGmax is highly correlated with ΔGtot for a large variety of modified surface structures, enabling the use of the Gibbs free energy of reaction as a descriptor in the optimization of the surface stability.
Ultimately, this study aims to advance sustainable energy technologies by developing robust and cost-effective catalysts for water electrolysis through knowledge-based proofing of the anode material against corrosion, facilitating the efficient use of renewable energy resources for green hydrogen production.
For surface computations, a (4 × 2 × 1) k-point mesh is employed for (2 × 2) superstructures of the (110) surface, with a constant k-point density in reciprocal space for larger surface configurations. These asymmetric slabs comprise five RuO2 trilayers, with two fixed and three fully relaxed trilayers. The relaxed bulk lattice parameters used in the calculations are optimized for RuO2, IrO2, and TiO2. Specifically, the lattice constants for RuO2 are a = b = 4.497 Å and c = 3.105 Å, for IrO2 they are a = b = 4.499 Å and c = 3.146 Å, and for TiO2 the parameters are a = b = 4.589 Å and c = 2.958 Å. A vacuum layer of 17.359 Å is employed. An energy cutoff of 450 eV and convergence criteria of 10−6 eV for energy during structural optimization is specified. Pseudopotentials are utilized for various elements including Ru, Ti, Ir, W, Re, Pt, Ta, O, and H19,21 designated as Ru_sv, Ti, Ir, W, Re_pv, Pt_pv, Ta_pv, O, H.
For Ti, a Hubbard coefficient UH of 8 eV is implemented in the Generalized Gradient Approximation (GGA) calculation to accurately describe the localized nature of Ti d-electrons in the semiconductor, thereby improving the characterization of its electronic structure.22–24 Other elements in the system did not require a UH correction, as their electron interactions were sufficiently well-described by conventional DFT.
The reactions investigated in this study include the oxygen evolution reaction at the anode, represented by the following reaction:
The corrosion reaction at the anode is:
RuO2 + 2H2O → RuO4 + 4H+ + 4e−. |
At the cathode, the hydrogen evolution reaction is represented by:
4H+ + 4e− → 2H2. |
Our computational approach is based on the total free energy equation, following methodologies detailed in previous studies.16,17,25 The free energy change for a particular configuration i, denoted as ΔGi(USHE,pH), is calculated using the following expression:
![]() | (1) |
The reason RuO4 was chosen in eqn (1) is that, at the potential corresponding to the dissolution threshold (around 1.21 V), RuO4 is typically the dominant solvated species. At higher potentials, ruthenium tends to stabilize in its higher oxidation state, often as RuO4. In this equation, ΔGi(USHE,pH) represents the free energy change for configuration i as a function of the applied potential relative to the standard hydrogen electrode (SHE), USHE, and the pH of the environment. The term EDFTi is the electronic energy of the system in configuration i, while EDFTref is reference energy, typically the energy of the initial state of the RuO2 system, obtained from density functional theory (DFT) calculations. The numbers nRuO4, ne−,i, nH2O,i, nH2,i, and nO2 represent the number of RuO4 units involved in the reaction or transformation for configuration i, the number of electrons transferred in the electrochemical process associated with configuration i, the number of adsorbed water molecules involved, the number of hydrogen molecules, and the number of oxygen molecules, respectively. The constant 0.059 eV represents a unit conversion factor at room temperature. USHE is the applied potential relative to the standard hydrogen electrode.
G RuO4,aq is the free energy of RuO4 in an aqueous solution, which can be determined using its entropy, SRuO4,aq ≈ 104.55 J (K−1 mol−1). GH2 and GO2 represent the free energies of hydrogen and oxygen molecules, respectively. The free energy of liquid water, GH2O,l, is determined by the expression
![]() | (2) |
At Ucorr, which is the thermodynamic corrosion potential where the system is in dynamic equilibrium, and the net current of oxidation (corrosion) and reduction reactions on the surface is zero, indicating no net electron flow. When pH = 0, indicating an acidic environment, the total free energy at the corrosion potential is given by:
ΔGi(Ucorr) = Ei − Eref + nRuO4GRuO4,aq − ne−,iUcorr − nH4OGH2O,l + nH2GH2 + nO2GO2 = 0 | (3) |
This equation is essential for understanding the stability and reactivity of RuO2 under different electrochemical conditions, providing insights into the factors influencing its performance as an oxygen evolution reaction (OER) catalyst.
Positive free energies indicate that corrosion is not thermodynamically favored, implying that the surface is thermodynamically stable against corrosion. The potential at which the free energy turns negative will hereafter be referred to as the corrosion potential (Ucorr), which indicates the lowest potential at which corrosion is thermodynamically preferred. Examining the corroded potential as a stability metric reveals variations in the stability of both atom types. The Rubr site involves the removal of one Rubr atom and one of its Obr atoms, while the other Obr atom relocates to the top of the nearest Rubr atom, which then becomes Ru5f. The corrosion potential of 1.277 V for the Rubr site is higher than the bulk corrosion potential of 1.21 V (the bulk corrosion potential refers to the total conversion of bulk RuO2 to RuO4), indicating that the site remains stable, as corrosion is not thermodynamically favored at this higher potential.
The Rubr–O site, formed by further removing the remaining O atom from the Rubr site (which previously moved to the nearest Rubr atom), transforms the nearest Rubr atom into a Ru4f atom without any termination on top. This site is identified as the most stable on the surface, with a corrosion potential of 1.488 V.
In contrast, the Rucus site is the least stable, indicating a higher susceptibility to corrosion. Its stability is comparable to that of bulk material, with a corrosion potential of 1.21 V. However, the instability of the Rucus site affects its neighboring Rucus atom, leading to the elongation of the underlying bond after metal dissolution, from 2.03 Å to 2.27 Å, as shown in Fig. 2b (top). In this specific case, our analysis reveals that the neighboring Rucus atom has a higher corrosion potential (1.291 V) compared to the initial Rucus site (1.214 V), making it less susceptible to dissolution.
We uncovered distinct dissolution behaviors and stabilities by extending our analyses to the (100) surface and step edge between the (100) and (110) facets. In each case, we examined the dissolution of different Ru sites with different oxygen terminations in the initial and final structures. Fig. 3 presents the summarized results, displaying the least stable site for each of the facets. Details for each of the facets can be found in Fig. S1–S3 in the ESI† for the (100), c(2 × 2)-reconstructed (100) and (100)/(110) step edge, respectively. For instance, Fig. S1 in the ESI† illustrates the dissolution behavior of the (100) surface, revealing the formation of a tetrahedral Ru4f structure post-dissolution and highlighting its comparatively lower stability: the least stable site on the (100) surface corrodes at a potential of 0.901 V, differing by 0.3 V compared to bulk RuO2.
Similarly, Fig. S2 in the ESI† elucidates the dissolution behavior of the (100)-c(2 × 2) surface, revealing corrosion potentials of 1.285 V for Ru4f (reported by Hess and Over)25 and 1.255 V for Ru6f (according to our investigation). Ru4f, Ru5f, and Ru6f denote distinct Ru crystallographic sites with fourfold, fivefold, and sixfold coordination, respectively, affecting the material's electronic structure and bonding.
Moving on to the dissolution behavior of step edge surfaces, Fig. S3 in the ESI† reveals corrosion potentials and stability characteristics. Notably, the step edge surface exhibits decreased stability compared to the bulk material, particularly evident in the least stable site (Rucus–O), with a corrosion potential of 0.858 V. The corrosion potentials for the least stable site on each facet are summarized in Fig. 3. This highlights the step edge as the least stable site among all examined, while the 110 surface emerges as one of the most stable surfaces. Consequently, we direct our further investigations towards the 110 surface.
![]() | ||
Fig. 4 (a and b) Influence of different dopant sublayers on RuO2 corrosion with a varied thickness (1L or 2L) of RuO2; (c) influence of different M0.21Ru0.79O2 (M = Ir, W) configurations on RuO2 corrosion (cf. Fig. S4†). Potentials were calculated at pH = 0. |
Furthermore, to distinguish the direct and indirect effect of the sublayer in RuO2 corrosion, we deposited 1 or 2 layers of RuO2 atop the dopant sublayer. In the case of 1 layer (1L), the corroded Ru atom is bonded to the sublayer via a single O3f (direct effect), while with 2 layers (2L), only an indirect effect of the sublayer is probed, since surface Ru is not directly bonded to the sublayer. This approach enables us to investigate whether varying the thickness of the RuO2 layer on the dopant sublayer affects the enhancement in stability against the RuO2(110) surface corrosion.
Fig. 4a and b illustrate the corrosion potential of the sublayer containing different dopants within the two models (1L, 2L). In the 1L model (direct effect), Ta, W, and Re destabilize the Ru top layer, whereas Ir, Pt, and Ti stabilize it against corrosion, as indicated by their decreased and increased corrosion potentials, respectively. The dopant is exposed after RuO4 dissolution, resulting in different surface stabilities of the final surface: W strongly prefers a higher oxidation state, making it the least supportive dopant for RuO2 surface stability against corrosion. Ta and Re prefer high oxidation states but not as strongly as W. In contrast, Ir, Pt, and Ti prefer lower oxidation states, with Ir and Ti being the most supportive dopants for RuO2 surface stability against corrosion The trends within the two groups persist in the 2L model, but the corrosion potentials for each of the dopants now scatter closely around the value for pure RuO2, with Ta, Ir, and Ti now causing a weak stabilization. For the others (W, Re, Pt), the effect on stability is negligible.
Continuing our investigation into the effect of doping on the stability of RuO2 against corrosion, we selected W from the group (Ta, W, Re) and Ir from the group (Ir, Pt, Ti), mixing them randomly with RuO2 to create eight configurations comprising a mixture of M0.21Ru0.79O2 (M = Ir, W). To maintain consistency and avoid lattice parameter issues, we fixed two layers of pure RuO2 at the bottom, while three layers of (M0.21Ru0.79O2 (M = Ir, W)) were allowed to relax (the considered dopant configurations are shown in Fig. S4†). Fig. 4c displays the corrosion potential of these configurations. In all configurations, Ir0.21Ru0.79O2 consistently demonstrates superior stability against corrosion compared to W0.21Ru0.79O2. The observed variations in stability among configurations are fairly consistent between the two dopants, suggesting that the positions of dopant atoms influence the corrosion behavior of the Rudiss atom, particularly its nearest neighbors. The distribution of dopant atoms in the bridge atom sites of the surface significantly influences stability, particularly the nearest bridge atom of Rudiss, as observed in configurations (3, 5, 7), where the bridge site next to Rudiss is occupied with a dopant atom. The metal atom directly underneath the Rudiss plays a similarly strong role: configuration 7 in Ir0.21Ru0.79O2, where Ir is the metal atom underlying Rudiss atom, is identified as the most stable configuration. Conversely, configuration 3 in W0.21Ru0.79O2, where Ru acts as the underlying Rudiss atom, demonstrates the highest stability among W0.21Ru0.79O2 configurations. This consistency aligns with sublayer results, where W and Ir in the sublayer destabilize and stabilize the surface Ru atom, respectively. This trend is equally observed in the random configurations. Conversely, configurations where the dopant is not located in the bridge position, such as configurations 6 and 8 in W0.21Ru0.79O2, which lack W bridge atoms, are the least stable. This analysis provides valuable insights into predicting the stability of the surface based on specific dopant sites.
Interestingly, while in the case of Ir, all configurations demonstrate a stabilization, W can provide both a stabilizing effect, if it is located directly at the surface, and a destabilizing effect, if it is located in the sublayer. The strong role of the dopant distribution is further emphasized if we further consider different dopant contents, only favoring the most stabilizing patterns identified from the above analysis: in this approach, we can achieve a corrosion potential of 1.457 V for an Ir content as low as 8% in the slab, only placing one Ir atom in the bridge position next to, and another directly underneath (cf. Fig. S5†). This represents a significant enhancement compared to the bulk corrosion of 1.21 V. By adding more Ir dopants in favorable positions, the corrosion potential can be further raised to 1.492 V at 21% Ir in the slab. This result highlights that a significant degree of stabilization can theoretically be achieved by doping very little Ir in just the right positions; however, experimentally achieving such selective doping remains a significant challenge. This result is significant because the price of raw Ir metal has been over 10 times as high as that of Ru consistently over the past three years.26 If Ir was used to stabilize RuO2-based anodes for OER, its content should be as low as possible to reduce the cost of raw materials. Our results presented herein enable a knowledge-based approach to achieve these goals, which remains to be tested by experimental model studies.
Using the lattice parameter of IrO2, Fig. 5a and b depict the corrosion potential of different thicknesses of RuO2 on two IrO2 substrate models consisting of either 3 or 4 layers (3L, 4L). We observe that there is a significant scatter in the Ucorr for even and uneven numbers of RuO2(110) layers. Similarly, we observe slightly different corrosion potentials for 3 and 4 layers of IrO2 substrate. Both models consistently show that odd layers of RuO2 stabilize the surface more strongly against even layers, and the corrosion potential converges around 6–7 layers to around 1.23 V for both models. According to these calculations, odd layers of RuO2 grown on the IrO2 clearly show a higher stabilization. Still, this behavior is likely significant only for very thin epitaxially grown layers not exceeding 4–5 layers. However, the lattice strain introduced due to IrO2 as a substrate led to an overall stabilization of the RuO2(110) surface by shifting the Ucorr from 1.21 V (pure RuO2(110)) to 1.23 V.
Such odd-even oscillations have been described previously for TiO2(110),27,28 and upon closer examination, we attribute it to the layer stacking in the (110) direction of the rutile lattice, which comprises two different types of atoms, Rucus and Rubr stacked in an A–B–A–B sequence on the substrate. For odd and even layers, corrosion occurs in an A and B layer, respectively, which may explain the observed oscillation, although the exact reason for this behavior is still unknown. This observation prompted us to reinvestigate the (100) surface, composed solely of A layers, as depicted in Fig. S6.† In this model, the oscillation of the Ucorr between odd and even layers is not observed. On a (100) surface, it is in steady observed that increasing thickness of strained RuO2 on top of IrO2, destabilizes the surface against corrosion, from 1.13 V (1L) to 0.84 V (7L). This indicates that strain can introduce a significant stabilizing effect in an otherwise inherently unstable facet like (100) and may be employed in addition to targeted doping to enhance the durability of RuO2-based OER catalysts. On the other hand, in the more stable (110) facet, the effect of strain is negligible if IrO2 is employed as a substrate.
To validate these conclusions, the corrosion potentials for RuO2 layers grown on top of IrO2 without strain (i.e., employing the RuO2 lattice parameter) are shown in Fig. 5c. Here, we observe the same odd-even oscillations; however, for thick RuO2 layers, the corrosion potential converges to the value of pure RuO2(110), confirming that the stability of the RuO2 surface is enhanced due to the strain introduced by the IrO2 substrate. The oscillations, on the other hand, are not a result of strain, but occur in thin RuO2(110) layers regardless of strain. We emphasize that, despite these oscillations, the strain induced by the IrO2 substrate enhances stability of RuO2(110), regardless of thickness.
For the TiO2 substrate, we utilized the same model as for the 3L IrO2 model. Fig. S7† depicts the corrosion potential of different thicknesses of RuO2 on a 3L TiO2 substrate with varying lattice parameters. The odd-even oscillation behavior occurs here as well, but with a lower intensity. Interestingly, when using the TiO2 lattice parameter, the lattice strain leads to destabilization. Conversely, aligning with the RuO2 lattice parameter results in slight stabilization, consistent with Section 3.1.2. The different stabilization behavior of IrO2 and TiO2 substrates is likely due to the different magnitude of lattice mismatch, but also due to the different orientations of strain and stress.
The investigation reveals that RuO2 layers grown on IrO2 substrates exhibit enhanced stabilization due to minimal lattice mismatch, while RuO2 on TiO2 substrates faces destabilization due to greater overall lattice mismatch, with odd-even oscillations in corrosion potential observed for both substrates in (110) direction. The (100) direction does not display such oscillations, but a significantly higher degree of stabilization due to the lattice mismatch.
For the removal of Rubr, we propose the following steps illustrated in Fig. 6, with the free energy profile depicted in Fig. 7:
![]() | ||
Fig. 6 The Rubr corrosion mechanism on the RuO2(110). Colors: sky blue for Rubulk, red for O, yellow for H, black for Rudiss (dissolution Ruthenium), and lime for Oads (adsorbed oxygen). |
![]() | ||
Fig. 7 The free energy profile of the Rubr corrosion mechanism on the RuO2(110) surface at an electrode potential USHE = 1.23 V. Free energies were calculated at pH = 0. |
Reconstruction step (a and b): this pivotal step initiates corrosion by lifting a Rubr atom, forming two new Ru–O bonds with the neighboring Oot. This Ru atom, now referred to as Rudiss, requires +1.180 eV to accomplish this first, energetically uphill step.
First H2O adsorption (c–g): in this step, an H2O molecule is initially adsorbed on top of the new undercoordinated Ru site. This step is followed by the desorption of two-electron (e−)–proton (H+) pairs. The process begins with H2O adsorption and progresses until only terminal oxygen remains. Consequently, the newly formed RuO6 species undergoes a second reconstruction, transitioning into a tetrahedral RuO4 species. During this transformation, one of the Obr atoms detaches from the underlying Ru, becoming another terminal O. Simultaneously, the Rudiss atom detaches from the underlying O3f and one of the neighboring Oot atoms. It is now connected to one Oot and one Obr, establishing two terminal O atoms in a four-fold coordination (Ru4f).
Second H2O adsorption (h and i): mirroring the preceding step, a second H2O molecule adsorbs. Upon the removal of the first H+/e− pair, the Rudiss atom severs its bond to the neighboring Obr, forming a RuO3(OH) complex attached only to Oot. The final removal of H+/e− is exergonic and leads to the formation of a RuO4 species.
Dissolution step (j and k): in the final step, the RuO4 detaches from the surface, leaving an empty Rucus site behind. Subsequently, this site is saturated by the adsorption of the final H2O molecule and the detachment of two H+/e− pairs.
The free energy profile of the Rubr corrosion mechanism on the RuO2(110) surface, as illustrated in Fig. 7, shows the free energy at each step. Notably, the initial reconstruction step (step b) presents the highest energetic threshold at 1.180 eV, indicating its unfavorable energetic characteristics. This chemical step is not influenced by the potential and likely represents the actual bottleneck under typical OER operating potentials.
Furthermore, removing the first H+/e− pair following the addition of the second H2O (step h) is also quite high in free energy at 1.120 eV. This is the electrochemical step in nature, and its free energy can be lowered by increasing the applied potential. This indicates that the initially formed RuO3(OH)/RuO4 complexes may have a significant lifetime on the surface and may act as possible active sites in the OER, as proposed by Klyukin et al.30
We examined the influence of various dopants on the reaction pathway toward RuO4 formation to gauge the potential enhancement in stability against RuO2(110) surface corrosion due to surface modification. Here, we distinguish between two possible modification strategies: (1) supporting RuO2 on a substrate of a different oxide, for which we only consider Ir, as it has a similar lattice parameter to RuO2, thereby avoiding strain effects, and (2) doping, i.e., the quasi-random substitution of Ru ions by other ions, and the respective influence on the reaction steps of the corrosion mechanism. Here, we consider Ti, Ir, and W, as representatives of the three groups favoring different oxidation states as identified in Section 3.1.2.
In Fig. 8, we illustrate the influence of the layer thickness on the free energy profile for the Rubr corrosion mechanism on the RuO2@IrO2(110) surface with two models (2L, 3L) of RuO2. We observed a significant correlation between the thermodynamic free energy of corrosion (represented by the free energy difference between the final state k and initial state a in Fig. 8) and reaction mechanisms, represented by the free energy of the highest intermediate (ΔGmax). The reaction steps of corrosion on the supported RuO2(110) surfaces largely follow the same sequence as on pure RuO2(110), only the energies are shifted: The corrosion intermediates on odd layers exhibit higher free energies than their even-layer counterparts, indicating increased stability and reduced susceptibility to corrosion, both in terms of thermodynamics and in terms of thermodynamic barriers. However, the highest energetic threshold, i.e., the reconstruction step (step b), remains unchanged for three layers of RuO2 (green line), standing at 1.197 eV. For two layers of RuO2 (orange line), the second H2O adsorption, coupled with the removal of a proton/electron pair (step h) requires a free energy of 1.01 eV, which is not significantly different from the initial reconstruction step (step b), indicating a possible shift in the reaction pathway at this potential (USHE = 1.23 V). However, at a potential typically applied during OER (USHE > 1.4 V), the initial reconstruction step, being a purely chemical step, remains the highest energetic threshold.
We initially focused on examining the Ti0.67Ru0.33O2 mixture, maintaining two fixed layers of RuO2 at the bottom while allowing three layers of Ti0.67Ru0.33O2 to relax. Fig. 9 focuses on the impact of Ti atom distribution on the free energy profile of the Rubr corrosion mechanism on the Ti0.67Ru0.33O2(110) surface. We consider two different configurations, each defining a specific neighborhood around the dissolving Rucus atom, to elucidate the influence on individual reaction steps. In configuration 1, the bridge atom next to the dissolving atom is replaced by Ti, while the neighboring cus atom participating in the dissolution mechanism remain Ru. In configuration 2, the neighboring bridge atom remains Ru, while the neighboring cus atom is replaced by Ti. We note that, on these doped surfaces, there are many possible realizations of the corrosion mechanism, involving the bridge and cus sites, and in some cases, different cus sites. We tested multiple configurations and bond-breaking sequences to identify the pathways with the lowest free energy. This comparative analysis, illustrated in Fig. S9,† highlights key differences between chosen and non-chosen configurations, revealing how dopants can alter the reaction mechanism compared to pure RuO2(110). The results shown below reflect the most favorable pathways identified.
This specific modification of the environment alters the energetics of the reaction steps as follows: configuration 1 (orange line in Fig. 9) exhibits higher free energy and greater stability compared to pure RuO2 (blue line) and configuration 2 (green line). The free energy reaches its maximum value of 2.075 eV in configuration 1 (orange line in Fig. 9) during the reconstruction step (step b), while configuration 2 reaches a maximum of 1.264 eV during the same step. Notably, in configuration 1, the nearest bridge neighbor to Rudiss is Tibr. During the reaction steps, especially in the second H2O adsorption step (step h), no proton/electron pair is removed, and the H adsorbs onto Obr (which is on Tibr). Then, in step j, the OH on Tibr is removed.
In configuration 2, the nearest neighbor is another Rubr. The transition from step a to step b, due to the unfavorable terminal O at Ticus in step a, involves three preparatory sub-steps (from b'′′ to b' shown in detail in Fig. S8†). These sub-steps include the adsorption of H2O on the Ticus atom. Upon the removal of the first H+/e− pair, the Rudiss atom severs its bond to the neighboring OH of Ticus. The final removal of H+/e− is exergonic, resulting in the formation of the regular step b. Notably, we observed a significant drop in free energies, attributed to changes in the reaction mechanism compared to pure RuO2(110), starting from step f. Unlike pure RuO2, where the tetrahedral RuO4 species is connected to neighboring Obr and Oot, in configuration 2, the RuO4 species prefers being connected to two Oot, resulting in an exergonic reaction step (e → f). This is mainly because it is highly unfavorable for Ticus to be capped by terminal oxygen; moving the RuO4 species to the cus sites avoids this scenario. Similarly, upon the removal of the final H+/e− pair in the second H2O adsorption step, the Rudiss atom severs its bond with the neighboring Oot of Rucus akin to the Walden inversion proposed by Hess et al.25 This results in the formation of a RuO4 species attached only to the Oot of Ticus, thereby avoiding the unfavorable terminal O at Ticus. In the final step, RuO4 detaches, leaving an empty Ticus site, as Ticus prefers termination without O.
Moving on to the Ir0.67Ru0.33O2 mixture, we adopted a similar methodology by fixing two layers of RuO2 at the substrate while allowing three layers of Ir0.67Ru0.33O2 to relax. The distribution of Ir atoms in configurations 1 and 2 of Ir0.67Ru0.33O2 mirrors that of Ti atoms in Ti0.67Ru0.33O2.
Fig. 10 illustrates how the Ir distribution influences the free energy profile of the Rubr corrosion mechanism on the Ir0.67Ru0.33O2(110) surface. Like in Fig. 9, we explore two configurations: configuration 1 (orange line) shows a ΔGmax of 2.248 eV during the dissolution step (step j), whereas configuration 2 (green line) peaks at 1.479 eV during the second H2O adsorption step (step h). Configuration 1 exhibits higher free energy and greater stability compared to pure RuO2 (blue line) and configuration 2. Notably, in configuration 1, the nearest neighbor to the Rudiss is Irbr. Starting from the second H2O step (step h), upon the removal of the first H+/e− pair, the Rudiss atom severs its bond with the neighboring Obr–Irbr, resulting in the Irbr being saturated by terminal OH. The subsequent steps take place exclusively on the cus sites. In configuration 2, the nearest neighbor is another Rubr, and no changes in reaction mechanisms occur after step f, unlike configuration 2 of Ti0.67Ru0.33O2.
Lastly, we investigated the W0.21Ru0.79O2 mixture, employing the same layer arrangement as in the previous mixtures. Similar to the previous examinations, we chose two configurations: configuration 1, where the neighboring bridge and cus atoms are W and Ru, respectively, and vice versa, in configuration 2. We selected a lower doping content in this case because W is the least similar to Ru, and addition of too much dopant may alter the properties in unpredictable ways; nevertheless, significant effects are observed here, because the specific doped sites have a larger effect on the corrosion behavior than the overall doping content, as shown in Section 3.1.2.
Fig. 11 depicts how the W distribution influences the free energy profile of the Rubr corrosion mechanism on the W0.21Ru0.79O2(110) surface. It explores two configurations relative to pure RuO2. Notably, configuration 2 (green line) shows the highest free energy step at 1.144 eV during the final dissolution step (step j), while configuration 1 (orange line) peaks at 1.182 eV during the first reconstruction (step b).
![]() | ||
Fig. 11 The free energy profile at USHE = 1.23 V illustrates the Rubr corrosion mechanism on the W0.21Ru0.79O2(110) surface for two configurations. Free energies were calculated at pH = 0. Colors: sky blue for Rubulk, red for O, black for Rudiss, and violet for W. Note: step j consists of three steps represented by the highest free energy among them; please refer to Fig. S10† for details. *. |
The dramatic change in configuration 2 can be explained by the strong preference of Wcus to remain terminated by O at the given potential: upon the removal of the final H+/e− pair in the second H2O step, the Rudiss atom severs its bond with the neighboring Obr of Ru, forming a RuO4 species attached only to the Oot of W. While in the regular mechanism considered for all other cases, RuO4 desorbs, along with the Oot, Wcus requires to remain saturated. Therefore, before the dissolution step (step j), to lower this high free energy, a third H2O molecule adsorbs, detaining two H+/e− pairs, as shown in Fig. S10.† This alteration in the mechanism akin to the Walden inversion proposed by Hess et al.25 reduces the highest free energy step from 1.305 eV to 1.144 eV.
From the different behavior of the three dopants and their different termination preferences we learn that the choice of dopant can alter not only the free energies of the reaction steps, but also the nature of these steps. To avoid states that are too high in free energy due to unfavorable terminations of the metal sites next to Rudiss, additional H2O molecules may partake in the reaction toward RuO4. On Ti-doped RuO2, we observe that the initial reconstruction is complicated by the lack of terminal O at Ticus; however, the final states involving the removal of RuO4 are simplified as they allow the Ticus to recover its vacant state. On the other hand, in W-doped RuO2, Wcus has such a strong preference to be capped with Ocus that additional steps are required late in the mechanism to detach RuO4 from the surface.
Close inspection of the data presented in Fig. 7–11 suggests that the thermodynamics of corrosion are highly correlated across a large variety of surface modifications. Based on all the data presented in this section, we quantitatively examined the correlation between the thermodynamic corrosion potential and the ΔGmax of the full reaction energy profiles to assess the suitability of ΔGtot as a descriptor for screening studies. As depicted in Fig. 12, our correlation analysis reveals a robust positive relationship (R2 = 0.96) between the total free energy (0 ≤ ΔGtot ≤ 1.5 eV) and the maximum free energy (1.0 eV ≤ ΔGmax ≤ 2.2 eV). This finding provides compelling evidence of the inherent connection between these parameters, highlighting a fundamental relationship governing material stability across diverse surface configurations across a large range of free energies. The maximum free energy (ΔGmax) is particularly significant for kinetic studies as it often represents the potential-determining step of a reaction, thus influencing the overall reaction rate and, in the case of corrosion, stability. Since ΔGtot is a highly accessible quantity, requiring only the computation of initial and final state of the surface, we propose it as a descriptor suitable for screening the efficacy of different surface modifications via computation. Furthermore, this suggests that the stepwise corrosion mechanism does not need to be known in order to predict relative stabilities between similar surfaces.
![]() | ||
Fig. 12 Correlation between ΔGtot and ΔGmax illustrates material stability interconnections across surfaces. Free energies were calculated at pH = 0 and USHE = 1.23 V. |
Overall, our investigation into Ti0.67Ru0.33O2, Ir0.67Ru0.33O2, and W0.21Ru0.79O2 mixtures unveils nuanced variations in reaction mechanisms and surface atom distributions. These findings underscore the critical role of dopant selection in optimizing RuO2 surface stability against corrosion. By elucidating the intricate interplay between dopants, surface terminations, and reaction pathways, our study provides valuable insights for the design and engineering of advanced materials exhibiting long-term stability for applications ranging from catalysis to energy storage.
Several studies have explored the incorporation of dopants into RuO2 to improve stability. Kasian et al.10 focused on the stabilization effect of Ir in RuO2, demonstrating that incorporating 25% Ir significantly enhances stability. Similarly, Escalera-López et al.34 investigated the electrochemical activity–stability relationship of Ir–Ru mixed oxides (IrxRu1−xO2), finding that 20% Ir improves stability while maintaining high activity. Both studies indicated a significant positive effect on surface stability with a slight decrease in activity, which informed our chosen Ir concentration in Fig. 4c. Our results are consistent with these findings, showing a significant improvement in the stability of the RuO2(110) surface. In studies exploring OER activity and electrocatalyst stability, IrxRu1−xOy catalysts with Ir concentrations up to 70% were developed, achieving 3 A cm−2 at 1.8 V with strong stability in PEM electrolysis, though high turnover frequency altered Lewis acidity, impacting water oxidation.35 Additionally, tuning the near-surface composition of Ru and Ir via surface segregation resulted in a Ru0.5Ir0.5 alloy exhibiting four times higher stability than the best Ru–Ir OER materials, without compromising catalytic activity.9 This is consistent with our investigation of the Ir0.67Ru0.33O2 mechanism, where the high Ir concentration contributes to one of the most stable surfaces, as shown in Fig. 10. We note that doping RuO2 with Ir poses an economic challenge, as the high cost of Ir may limit practical application. Our findings indicate that substituting specific Ru sites by Ir may result in significant enhancement, an insight that could be exploited to reduce the Ir content of mixed (Ru,Ir)O2 for OER applications.
Our study investigates the influence of lower cost doping elements, such as Ti and W, which have been proposed in the literature as potential dopants for RuO2. Hao et al.12 investigated the stability and activity of RuO2 with W and Er (W0.2Er0.1Ru0.7O2−δ). They found that adding W and Er significantly increases the catalyst's oxygen vacancy formation energy, improving durability and activity. This finding informed our chosen W concentration in Fig. 4c and 11 of the reaction mechanism investigation. Our predictions in Fig. 4c allow us to identify optimal dopant sites that enhance the durability and stability of the M0.21Ru0.79O2(110) surface.
For TixRu1−xO2, Godínez-Salomón et al.11 demonstrated that incorporating Ti (12.5–50%) enhances catalyst stability and reduces Ru dissolution, with low Ti concentrations (12.5–20 at%) significantly improving stability compared to pure RuO2. Regarding IrO2 substrates, most studies focus on RuO2 substrates enhancing the surface activity of IrO2. Our study uniquely examines the effect of RuO2 thickness on an IrO2 substrate, identifying the optimal thickness for maximal stability. Fig. 5 reveals a distinct odd-even layer behavior that suggests that uneven layers, such as 1 or 3, exhibit the highest stability.
Consistent with recent findings by Chaudhary et al.,36 our study shows that tensile strain (as seen in RuO2 on IrO2) enhances corrosion resistance, while compressive strain (as observed in RuO2 on TiO2) leads to destabilization of the surface, highlighting the critical role of strain in influencing the corrosion behavior of RuO2.
In terms of reaction mechanism, while many studies have concentrated on the oxygen evolution reaction (OER) mechanism at RuO2(110), fewer have delved into the specifics of RuO2 corrosion mechanisms. Klyukin et al.30 proposed a concept wherein a defect-free surface undergoes transitions leading to RuO4 formation and subsequent corrosion. Similarly, Liu et al.29 introduced four potential initial reconstruction steps (chemical steps), focusing on one to elucidate a mechanism that results in RuO4 formation and surface corrosion. We adopted and expanded upon this mechanism in our investigation.
Our study extends beyond merely detailing the corrosion mechanisms of the RuO2 surface to examine the impact of doping on these mechanisms, as shown in Fig. 8–11. Our findings highlight how the termination of the cus atoms in dopants affects the reaction mechanism and the associated free energies. In contrast to Gong et al.,37 who use metal–oxygen bond strength from MD simulations to predict dissolution rates of complex glasses, our study focuses on thermodynamic descriptors, specifically ΔGmax and ΔGtot, to assess corrosion resistance in RuO2. While Gong et al.37 report R2 values of 0.80–0.92 for bond strength and dissolution rates, our study finds a strong correlation (R2 = 0.96) between ΔGmax and ΔGtot, demonstrating their effectiveness in predicting material stability. Similarly, Schwöebel et al.38 use hydrogen bond acceptor strength as a descriptor for organic and inorganic compounds, demonstrating a high correlation (R2 = 0.97) with experimental data, while our study employs thermodynamic parameters to predict corrosion stability, highlighting the broad applicability of descriptor-based methods across different material types. The concept of 'maximum free energy (ΔGmax)' as a descriptor for stability was first introduced by Exner and Over,39 and our findings further support its utility in understanding reaction rates and mechanisms the trends of reaction rates and mechanisms upon altering the properties of the surface. Further work is required to understand the details of the kinetic processes governing the competition between the OER and the parasitic corrosion leading toward electrode dissolution. Kinetic studies are crucial for offering insights into activation energies, rate constants, and the effects of temperature and other reaction conditions.40–51 Additionally, we plan to investigate the OER activity of various doped surfaces to compare their performance and stability relative to pure RuO2.
Our findings emphasize the critical role of understanding the intricate interplay between surface composition, dopants, and reaction mechanisms to optimize the stability of RuO2 surfaces. Dopants such as Ir, Ti, and W exhibit diverse effects on stability, with Ir demonstrating superior stabilization compared to other elements. Specifically, Ti favors termination without O, Ir favors OH termination, and W favors O termination, each influencing distinct free energy profiles and reaction pathways, particularly evident in configuration 2 across all mixtures. Based on the correlation between the total change in Gibbs free energy (ΔGtot) and the maximum Gibbs free energy (ΔGmax) we propose ΔGtot as an accessible descriptor to predict the effect of different surface modifications on the stability of surface sites.
Overall, our study not only deepens our understanding of the complex behavior of RuO2 surfaces but also offers valuable insights into designing tailored materials with enhanced stability for various applications, including catalysis and energy storage. These findings lay a solid foundation for future research aimed at optimizing the performance of RuO2-based materials in real-world applications, thereby advancing the field of materials science and engineering.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta00218d |
This journal is © The Royal Society of Chemistry 2025 |