2,753
Views
15
CrossRef citations to date
0
Altmetric
Research Article

Hydrogen oxidation on oxygen-rich IrO2(110)

, , ORCID Icon, ORCID Icon &
Pages 1-13 | Received 18 Sep 2018, Accepted 21 Dec 2018, Published online: 31 Jan 2019

ABSTRACT

We investigated the adsorption and oxidation of H2 on O-rich IrO2(110) using temperature programmed reaction spectroscopy (TPRS) and density functional theory (DFT) calculations. Our results show that H2 dissociation occurs efficiently on O-rich IrO2(110) at low temperature and initiates from an adsorbed H2 σ-complex on the coordinatively-unsaturated Ir atoms (Ircus). We find that on-top oxygen atoms (Oot), adsorbed on the Ircus sites, promote the desorption-limited evolution of H2O during subsequent oxidation of the adsorbed hydrogen on IrO2(110) while suppressing reaction-limited production of H2O via the recombination of bridging HO groups (HObr) (~500 to 750 K) during TPRS. The desorption-limited TPRS peak of H2O shifts from ~490 to 550 K with increasing Oot coverage, demonstrating that Oot atoms stabilize adsorbed OH and H2O species. DFT predicts that molecularly-adsorbed H2 dissociates on O-rich IrO2(110) at low temperature and that the resulting H-atoms redistribute to produce a mixture of HObr and HOot groups, with equilibrium favouring HOot groups. Our calculations further predict that subsequent H2O evolution occurs through the recombination of HObr/HOot and HOot/HOot pairs, and that these reactions represent desorption-limited pathways because the dissociative chemisorption of H2O is favoured over molecular adsorption on IrO2(110). The higher stability of HOot groups and their preferred formation causes the higher-barrier HOot/HOot recombination reaction to become the dominant pathway for H2O formation with increasing Oot coverage, consistent with the experimentally-observed upshift in the H2O TPRS peak temperature.

Graphical abstract

Introduction

Understanding the interactions of hydrogen with IrO2 surfaces is important for improving applications of electrocatalytic water splitting and developing IrO2-based catalysts that can efficiently convert light alkanes to value-added products. Prior studies show that IrO2 exhibits high activity towards the electrocatalytic conversion of water to hydrogen and oxygen [Citation1-Citation4], and thus motivate fundamental investigations of the chemical properties of well-defined IrO2 surfaces towards species that are involved in water-splitting chemistry (e.g., H2, O2, H2O). Recent studies also demonstrate that the IrO2(110) surface efficiently promotes the C-H bond cleavage of methane, ethane and propane at low temperatures (<150 K) [Citation5-Citation7], and that the resulting alkyl groups react at higher temperature during temperature-programmed reaction spectroscopy (TPRS). Our prior results show that surface H-atoms can have a strong influence on the chemistry of alkyl groups on IrO2(110), and thus further motivate efforts to clarify the fundamental interactions between hydrogen and IrO2 surfaces.

Late transition-metal oxides can be highly active for oxidizing H2, hydrocarbons and other compounds, provided that the surfaces expose pairs of coordinatively-unsaturated (cus) metal and oxygen atoms [Citation8,Citation9]. The stoichiometric termination of rutile RuO2(110) and IrO2(110) surfaces exposes equal quantities of cus-metal and oxygen atoms, where the cus-oxygen atoms are referred to as bridging O-atoms (Obr) because each Obr atom occupies a bridging site relative to two underlying metal atoms in the oxide lattice. Oxygen atoms can also bond directly on-top of the cus-metal atoms to produce a singly-coordinate species referred to as an on-top oxygen atom (Oot). Prior studies show that O2 dissociation occurs efficiently on RuO2(110) at room temperature, and can generate variable coverages of Oot atoms that coexist with Rucus and Obr atoms [Citation8-Citation10]. These studies reveal that both Obr and Oot atoms can be highly reactive, but that Oot atoms are generally more active as H-atom acceptors. Studies to clarify the reactivity of Obr vs. Oot atoms are important for developing atomic-level descriptions of oxidation catalysis promoted by IrO2(110).

Past investigations demonstrate that stoichiometric (s-) RuO2(110) is reactive for oxidizing H2 and that Oot atoms significantly enhance the surface activity [Citation11-Citation17]. At low temperature, hydrogen dissociation initiates from H2 molecularly-adsorbed on Rucus atoms, where the H2 molecule datively bonds with a Rucus atom to produce an adsorbed H2 σ-complex [Citation11,Citation16-Citation19]. Experiments show that H2O evolves in a reaction-limited TPRS feature (~500 to 700 K) during H2 oxidation on s-RuO2(110), such that H2O forms at temperatures higher than that required for chemisorbed H2O to desorb from the surface. The rate-limiting step in H2 oxidation on s-RuO2(110) has been attributed to the decomposition of stable HObr groups [Citation12]. In contrast, Oot-atoms promote the desorption-limited evolution of H2O (~400 K) during H2 oxidation on RuO2(110) [Citation17], demonstrating that the Oot atoms provide low-energy pathways for H2O formation. Knapp et al. report that Oot atoms efficiently harvest H-atoms from Obr atoms, resulting in the formation of chemisorbed H2Oot species at low temperature [Citation13-Citation15]. We have recently found that H2 adsorption and oxidation occurs on s-IrO2(110) by a similar mechanism as reported for s-RuO2(110) [Citation20]. A key difference is that H2 interacts more strongly with Ircus compared with Rucus atoms, resulting in a higher molecular H2-binding energy on IrO2(110) and facile H2 dissociation at 90 K. After H2 dissociation, H2O evolves from IrO2(110) in a reaction-limited TPRS feature (~400–750 K) attributed to the recombination of HObr groups.

In the present study, we investigated H2 oxidation on O-rich IrO2(110) using TPRS and DFT calculations. We find that Oot atoms on IrO2(110) promote H2 oxidation during TPRS by facilitating desorption-limited H2O evolution but that Oot atoms also stabilize HOot and H2Oot species, resulting in higher barriers for H2O desorption from O-rich vs. stoichiometric IrO2(110). We present evidence that H-atoms transfer rapidly from Obr to Oot atoms on IrO2(110) and that the desorption-limited evolution of H2O occurs through recombination of HOot/HObr as well as HOot/HOot pairs, with the latter pathway becoming increasingly important with increasing Oot coverage.

Experimental details

Details of the UHV analysis chamber with an isolatable ambient-pressure reaction cell utilized in the present study have been reported previously [Citation5]. Briefly, the Ir(100) crystal employed in this study is a circular disk (9 mm × 1 mm) that is attached to a liquid-nitrogen-cooled, copper sample holder by 0.015” W wires that are secured to the edge of crystal. A type K thermocouple was spot welded to the backside of the crystal for temperature measurements. Resistive heating, controlled using a PID controller that varies the output of a programmable DC power supply, supports linearly ramping from 80 to 1500 K and maintaining the sample temperature. Sample cleaning consisted of cycles of Ar+ sputtering (2000 eV, 1.5 μA) at 1000 K, followed by annealing at 1500 K for several minutes. The sample was subsequently exposed to 5 × 10−7 Torr of O2 at 900 K for several minutes to remove surface carbon, followed by flashing to 1500 K to remove final traces of oxygen. We considered the Ir(100) sample to be clean when we obtained sharp low-energy electron diffraction (LEED) patterns consistent with the reconstructed (5 × 1) structure and did not detect impurities using AES and detected negligible CO and CO2 production during flash desorption after adsorbing oxygen.

We generated an IrO2(110) film by exposing Ir(100) to 5 Torr O2 (Airgas, 99.999%) for a duration of 10 min (3 × 109 Langmuir) in the ambient-pressure reaction cell at a surface temperature of 765 K. Our ambient-pressure reaction cell is designed to reach elevated gas pressure while maintaining UHV in the analysis chamber [Citation5]. After preparation of the oxide film, we lowered the surface temperature to 600 K, and then evacuated O2 from the reaction cell and transferred the sample back to the UHV analysis chamber. We exposed the film to ~23 L of O2 while cycling the surface temperature between 300 and 650 K to fill oxygen vacancies that may have been created during sample transfer from the reaction cell to the analysis chamber. This procedure produces a high-quality IrO2(110) surface that has a stoichiometric surface termination and consists of ~10 layers of IrO2(110), corresponding to a thickness of 3.2 nm.

shows a ball and stick model of IrO2(110) partially covered with Oot atoms. The IrO2(110) surface unit cell is rectangular, with bulk-terminated dimensions of a = 3.16 Å and b = 6.36 Å, and the surface consists of alternating rows of Ircus and Obr atoms along the [001] direction. Each of these surface species has a single dangling bond due to the decrease in bond coordination relative to bulk IrO2; the Ircus atoms have 5-fold coordination whereas bulk Ir atoms have 6-fold coordination, and Obr atoms have 2-fold coordination whereas bulk O-atoms have 3-fold coordination. One may consider that formation of the stoichiometric termination of IrO2(110) involves cleaving equal numbers of Ir-O and O-Ir bonds, and thus generating equal quantities of Ircus and Obr atoms at the surface. On the basis of the IrO2(110) unit cell, the areal density of Ircus atoms and Obr atoms is equal to 37% of the Ir(100) surface atom density of 1.36 × 1015 cm2. Since Ircus atoms are active adsorption sites, we define 1 ML as equal to the density of Ircus atoms on the IrO2(110) surface. On-top oxygen atoms (Oot) bond directly on Ircus atoms and also expose a dangling bond perpendicular to the surface (). Similar to the behaviour of oxygen on RuO2(110), TPD measurements show that Oot atoms on IrO2(110) are less stable that Obr atoms with Oot atoms desorbing between 400 and 650 K and Obr atoms desorbing in a TPD peak near 950 K [Citation5,Citation20].

Figure 1. Model representation of top and side views of the IrO2(110) structure with an Oot atom. The Ircus, Ir6f, Obr, Oot and O3f atoms are indicated, where Ir6f and O3f corresponds to 6-fold and 3-fold coordination. Red and blue atoms represent O and Ir atoms, respectively, and the Oot atom is shown in orange.

Figure 1. Model representation of top and side views of the IrO2(110) structure with an Oot atom. The Ircus, Ir6f, Obr, Oot and O3f atoms are indicated, where Ir6f and O3f corresponds to 6-fold and 3-fold coordination. Red and blue atoms represent O and Ir atoms, respectively, and the Oot atom is shown in orange.

We investigated the adsorption and reactivity of H2 (Airgas, 99.999%) on Oot-covered (‘O-rich’) IrO2(110) using TPRS. After a hydrogen exposure, we positioned the sample in front of a shielded mass spectrometer at a distance of ~5 mm and heated at a constant rate of 1 K/s until reaching a sample temperature of 800 K. Because the IrO2(110) starts to decompose above 650 K, it was necessary to prepare a fresh IrO2(110) film for each adsorption/reaction experiment. Reproducibility in our TPRS results provides evidence that we can repeatedly generate IrO2(110) films with nominally the same surface structure and composition. We quantified adsorbate coverages and desorption yields from TPRS using established procedures as described in the SI.

Computational details

All plane wave DFT calculations were performed using the projector augmented wave pseudopotentials [Citation21] provided in the Vienna ab initio simulation package (VASP) [Citation22,Citation23]. The Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional [Citation24] was used with a plane wave expansion cutoff of 400 eV. We used four layers to model the IrO2(110) film which has an ~12 Å thick slab. The PBE bulk lattice constant of IrO2 (a = 4.54 Å and c = 3.19 Å) is used to fix the lateral dimensions of the slab. The bottom two layers are fixed, but all other lattice atoms are allowed to relax during the calculations until the forces are less than 0.05 eV/Å. A vacuum spacing of ~25 Å was included, which is sufficient to reduce the periodic interaction in the surface normal direction. In terms of system size, a 2×4 unit cell with a corresponding 2×2×1 Monkhorst-Pack k-point mesh is employed. We also have performed select tests of H2, O2 and H2O adsorption energies with a large k-point mesh of 4 × 4 × 1. The difference with the larger k-point mesh is less than ~ 0.1 kJ/mol. Unless otherwise noted, our DFT calculations were performed for a single H2 molecule adsorbed within the 2×4 surface model of IrO2(110), and corresponds to an H2 coverage equal to 12.5% of the total density of Ircus atoms and 25% of the Ircus density within one Ircus row. In the present study, we define the binding energy, Eb, of an adsorbed H2 molecule on the surface using the expression,

Eb=EH2+EsurfEH2/surf

where EH2/surf is the energy of the state containing the adsorbed H2 molecule, Esurf is the energy of the bare surface, and EH2 is the energy of an isolated H2 molecule in the gas phase. From the equation above, a large positive value for the binding energy indicates a high stability of the adsorbed H2 molecule under consideration. We evaluated the barriers for H2 oxidation on the oxygen-rich IrO2(110) surface using the climbing nudged elastic band (cNEB) method [Citation25]. All the activation energy barriers that we report here have an imaginary vibrational frequency at the transition state. All reported binding energies and energy barriers are corrected for zero-point vibrational energy.

Experimental results

Chemisorption of O2 on IrO2(110) film at 85 K versus 300 K

,) show O2 TPD spectra obtained after adsorbing various quantities of oxygen on IrO2(110) at temperatures of 85 K and 300 K, respectively. Total oxygen coverages (atomic + molecular species) are specified in and given in units of ML of O-atoms. After saturating IrO2(110) with oxygen at 85 K, the O2 TPD spectrum exhibits two main features, one below 200 K and the other between ~400 and 650 K (). We have previously assigned these low and high temperature TPD features to the desorption of molecularly-adsorbed O2 and the recombinative desorption of Oot atoms on the Ircus rows, respectively [Citation20]. Analogous O2 desorption behaviour from RuO2(110) supports these assignments [Citation8-Citation10]. Furthermore, we find that a maximum of 0.86 ML (O-atom basis) of O2 desorbs in the features below 650 K, in agreement with the jamming coverage predicted by the random sequential adsorption (RSA) model for immobile dimers adsorbing on a linear array of sites. We note that an average of 0.11 ± 0.02 ML of H2O desorbs from the O2-exposed surfaces due to the adsorption of H2 from the vacuum background before or during the O2 exposures (Supporting Information, Figure S1). Initial TPRS experiments with 18O2-exposed IrO2(110) show that hydrogen reacts nearly exclusively with on-top oxygen when hydrogen is the limiting reactant. Thus, since the background hydrogen uptake is less than the lowest Oot coverage that we studied, it is reasonable to assume that the initial Oot coverage is equal to the sum of the H2O and O2 desorption yields below 650 K. Here, we describe the evolution of the O2 TPD spectra with increasing oxygen coverage on IrO2(110) and compare the behaviour with that reported previously for oxygen adsorbed on RuO2(110).

Figure 2. O2 TPD spectra obtained after exposing a s-IrO2(110) film to varying quantities of O2 at (a) 85 K and (b) 300 K.

Figure 2. O2 TPD spectra obtained after exposing a s-IrO2(110) film to varying quantities of O2 at (a) 85 K and (b) 300 K.

After adsorption at 85 K, O2 initially desorbs in a TPD feature (δ2) centred at ~570 K that broadens and shifts to ~550 K as the on-top oxygen coverage initially increases (). A second feature (δ1) develops and downshifts from ~500 to 470 K with increasing oxygen coverage above ~0.30 ML. The broadening of the O2 TPD feature towards lower temperatures suggests that neighbouring Oot atoms interact repulsively, and thereby lower the energy barrier for recombinative Oot desorption. Prior studies also report strong repulsive interactions between Oot atoms on RuO2(110) [Citation26,Citation27], and a significant broadening of the O2 TPD feature with increasing Oot coverage [Citation28]. The total desorption yield in the δ1 + δ2 feature is equal to 0.46 ML at saturation of the on-top oxygen layer at 85 K. Adsorption of O2 at 300 K also produces a δ TPD feature that saturates at an Oot coverage of 0.44 ML ().

Molecular oxygen begins to desorb in several, overlapping TPD features (γ1 – γ4) below ~300 K, with the features developing sharply as the oxygen coverage increases above ~0.35 ML at 85 K after the δ1 feature first appears. We label the O2 TPD maxima at ~102, 125, 150 and 180 K as γ1, γ2, γ3 and γ4 and note that these features populate nearly uniformly when they first become evident but that the γ2 peak at 125 K becomes dominant as the coverage increases towards saturation. The small shoulder on the trailing edge of the γ4 peak may arise from O2 adsorbed on defect sites or a minority phase within the oxide layer. The appearance of multiple, low-temperature TPD features may indicate that molecularly-adsorbed O2 adopts several binding geometries and that the binding energies are also sensitive to the local environment at the high oxygen coverages at which these species exist on IrO2(110). Our results demonstrate that O2 dissociation to produce Oot atoms is facile on the IrO2(110) surface. This behaviour is analogous to O2 on RuO2(110); however, an important difference is that O2 dissociation produces a lower saturation coverage of Oot atoms on IrO2(110) compared with RuO2(110) (~0.45 ML vs. 0.86 ML) in UHV. An implication is that molecular adsorption becomes favoured over O2 dissociation at lower Oot coverage on IrO2(110) relative to RuO2(110). Also, molecularly adsorbed O2 desorbs in a well-defined peak from RuO2(110) [Citation10], whereas molecular O2 desorbs in multiple TPD features from IrO2(110). This difference is likely a consequence of the larger amount of molecularly-adsorbed O2 that desorbs from IrO2(110) rather than dissociating.

Lastly, the appearance of the low temperature O2 TPD features (<~ 200 K) mainly after the recombinative O2 TPD feature (~400 to 600 K) saturates demonstrates that molecularly adsorbed O2 predominantly dissociates on IrO2(110) at coverages up to ~0.35 ML and may further indicate that the molecularly adsorbed species dissociate at temperatures below ~200 K, i.e., below the temperatures at which O2(ad) species desorb. Our DFT calculations support this conclusion as they predict a barrier of only 14 kJ/mol for O2(ad) to dissociate on IrO2(110) for an O2 coverage of 0.25 ML (O-atom basis). A possible implication is that the desorption of molecularly adsorbed O2 occurs only when Oot atoms are also present in large amounts on the Ircus rows. Further study is needed to examine how Oot atoms influence the binding of O2(ad) species on the IrO2(110) surface.

TPRS spectra of H2 oxidation on O-rich IrO2(110)

shows H2 and H2O TPRS traces obtained after exposing a saturation amount of H2 to IrO2(110) surfaces held at 85 K and initially covered with different quantities of on-top oxygen. We prepared the O-rich IrO2(110) surfaces by first saturating with O2 at 85 K, followed by flashing the surface to temperatures selected to desorb a specific fraction of the on-top oxygen. Hydrogen oxidation on stoichiometric-IrO2(110) produces H2O over a wide temperature range that can be decomposed into a TPRS peak near 490 K (α1) and a broad feature between ~500 and 750 K (α2). We have previously attributed the α1 and α2 features to the desorption-limited evolution of H2O (see Figure S2) and the reaction-limited recombination of HObr groups, respectively [Citation20]. At saturation of the H2 layer on s-IrO2(110), the H2 TPRS spectrum also exhibits small peaks at ~200 and 530 K that have been assigned to H2 desorbing from a molecularly-adsorbed σ-complex state and the recombination of H-atoms, respectively. We estimate an H2 saturation coverage of ~0.68 ML on s-IrO2(110) at 85 K and find that more than 90% of the initially adsorbed H2 σ-complexes dissociate during TPRS with the majority oxidizing to H2O.

Figure 3. (a) H2 and H2O TPRS spectra obtained after exposing a saturation amount (0.5 L) of H2 to IrO2(110) surfaces held at 85 K and initially covered with different quantities of on-top oxygen. (b) O2 TPD spectra obtained from initial (solid) and H2-saturated (dashed) IrO2(110) surfaces with initial on-top oxygen coverages of 0.42 ML (blue) and 0.65 ML (orange). (c) Initial H2 coverage (black), H2O TPRS yields (red) and H2 TPRS yields (green) as a function of the initial Oot coverage.

Figure 3. (a) H2 and H2O TPRS spectra obtained after exposing a saturation amount (0.5 L) of H2 to IrO2(110) surfaces held at 85 K and initially covered with different quantities of on-top oxygen. (b) O2 TPD spectra obtained from initial (solid) and H2-saturated (dashed) IrO2(110) surfaces with initial on-top oxygen coverages of 0.42 ML (blue) and 0.65 ML (orange). (c) Initial H2 coverage (black), H2O TPRS yields (red) and H2 TPRS yields (green) as a function of the initial Oot coverage.

Our TPRS results show that on-top oxygen species promote the desorption-limited evolution of H2O (α1) during H2 oxidation on IrO2(110) while suppressing reaction-limited H2O production (α2). This finding demonstrates that on-top oxygen species provide low-energy pathways for the oxidation of H2 to adsorbed H2O or HO species on IrO2(110) such that these adsorbed species form at temperatures lower than that at which adsorbed water desorbs during TPRS. Increasing the initial Oot coverage to 0.10 ML causes the H2 TPRS peaks to nearly vanish while the desorption-limited H2O peak at 490 K intensifies proportionally and the reaction-limited H2O feature remains at about the same intensity as observed during H2 oxidation on s-IrO2(110). The desorption-limited H2O peak continues to intensify as the on-top oxygen coverage increases from 0.10 to ~0.40 ML while the reaction-limited H2O feature concurrently diminishes to a negligible level. The desorption-limited H2O peak also shifts from 490 to ~540 K with increasing Oot-coverage to about 0.40 ML, and the H2 TPRS peaks completely vanish once the initial Oot coverage surpasses ~0.20 ML. The enhancement in the desorption-limited H2O yield shows that Oot atoms on IrO2(110) promote H2 oxidation to adsorbed H2O or HO below the temperatures at which H2O desorption becomes appreciable during TPD. Further, the substantial upshift in the H2O peak temperature suggests that increasing quantities of the adsorbed H2O or HO products are stabilized as the initial Oot coverage increases. The desorption-limited H2O peak upshifts to 550 K but also diminishes sharply as the initial Oot coverage increases above 0.40 ML.

shows O2 TPD traces obtained from initial vs. H2-exposed IrO2(110) surfaces with initial on-top oxygen coverages of 0.42 and 0.65 ML. We exposed each O-rich IrO2(110) surface to a dose of H2 (0.5 L) that is sufficient to saturate the s-IrO2 (110) surface with H2 at 90 K, and note that the H2 saturation coverage decreases with increasing initial Oot coverage due to site blocking, as discussed further below. Comparison of the O2 TPD traces (solid vs. dashed) reveals that the leading-edge component (δ1) of the Oot-atom recombination feature preferentially diminishes during H2 oxidation. This behaviour suggests that partial hydrogenation of Oot-covered IrO2(110) reduces the population of Oot-atoms that are located in the interior of Oot-atom chains where they are destabilized by neighbouring Oot atoms. Below, we report DFT calculations which predict that the hydrogenation of Oot atoms to HOot groups is thermodynamically favourable and could thus eliminate repulsive interactions among neighbouring Oot atoms. While the O2 TPD traces provide clear evidence that Oot atoms are reactive towards H2, they are inconclusive about the reactivity of O2,ot because the traces exhibit only small decreases in the molecular O2 desorption yields after H2 adsorption. The O2 TPD data also demonstrate that a smaller quantity of H2 adsorbs and oxidizes for initial on-top oxygen coverage of 0.65 ML compared with 0.42 ML. This behaviour is expected because H2 dissociation requires vacant Ircus sites and the coverage of such sites decreases as the initial coverage of on-top oxygen species increases. We note, however, that the H2 coverage decreases more sharply than expected as the on-top oxygen coverage increases above ~0.4 ML. We discuss the sharp decrease in H2 uptake in more detail below.

shows the initial H2 coverage and the H2O and H2 desorption yields obtained during TPRS as a function of the initial on-top oxygen coverage on IrO2(110), where the initial H2 coverage is taken as the sum of the H2O and H2 TPRS yields. The figure also shows the temperatures at which the surface was heated during preparation of the initial on-top oxygen layer. The yield data shows that all the adsorbed H2 oxidizes to H2O during TPRS when the initial Oot-coverage is greater than ~0.10 ML and the H2 coverage is saturated. The initial H2 coverage first decreases slowly but exhibits an abrupt decrease at an Oot coverage of ~0.4 ML before settling at a value of ~0.16 ML at saturation of the on-top oxygen layer. The H2 coverage is expected to decrease in proportion to the coverage of Oot atoms since these species occupy Ircus sites and prevent H2 adsorption and dissociation. Molecularly-adsorbed O2,ot species may also be effective in blocking H2 adsorption, considering that similar amounts of molecular O2 desorb during TPRS performed with and without adsorbed H2 ().

We attribute the abrupt decrease in H2 uptake at an Oot coverage of ~0.4 ML to the dissociative adsorption of background H2 before preparation of the on-top oxygen layer and the effect of preparation temperature on the site occupancy of the resulting H-atoms. Background H2 adsorption may also cause the H2 coverage to settle at a value of 0.16 ML at saturation of the on-top oxygen layer, rather than decreasing further (). We find that an average of 0.16 ± 0.03 ML of H2 dissociatively adsorbs on IrO2(110) from the vacuum background when the on-top oxygen layer is prepared by heating an O2-saturated surface to selected temperatures. The amount of H2 uptake from the background is consistent with an initial H2 dissociative adsorption probability near unity. Our prior study shows that H2 dissociates efficiently to produce Hot atoms and HObr groups on s-IrO2(110) at 85 K, whereas the subsequent hopping of Hot atoms onto Oot or Obr atoms involves a higher energy barrier of 85 to 90 kJ/mol according to DFT. This prediction suggests that Hot atoms will occupy Ircus sites until the surface reaches a sufficient temperature (~350–400 K) to promote Hot hopping to O-atoms. Consistent with this idea, we have previously presented evidence that saturating s-IrO2(110) with H2 at 85 K, followed by heating to 380 K enables the surface to accommodate additional H2 at 85 K because Hot atoms transfer to Obr sites and thereby vacate Ircus sites needed for H2 adsorption [Citation20]. In the present study, we find that the abrupt decrease in H2 coverage near [Oot] ~ 0.40 ML coincides with a decrease in the Oot-preparation temperature below about 450 K (). This sudden decrease in the H2 uptake likely occurs because the rate of H-transfer from Ircus to O atoms decreases abruptly below ~450 K, resulting in a proportional decrease in the coverage of Ircus sites available for H2 adsorption at 85 K. In support of this interpretation, we find that the H2 uptake decreases at a more uniform rate with increasing Oot coverage in TPRS experiments for which we prepared the Oot layer by performing variable O2 exposures at 300 K (Figures S3 and S4). Overall, the H2 uptake behaviour provides further evidence that H2 oxidation on O-rich IrO2(110) initiates from the H2 σ-complex state and thus requires vacant Ircus sites to proceed at the conditions studied.

Figure 4. Images showing computed structures of molecular and dissociated H2O chemisorbed on IrO2(110) and binding energies defined relative to the specified initial states. (a) H2Oot species and (b) a HObr/HOot pair resulting from molecular vs. dissociative chemisorption of an H2O molecule on s-IrO2(110). (c) HObr/HOot-H2Oot dimer and (d) HObr/HOot + HObr/HOot dimer with binding energies per H2O defined relative to 2H2O(g) + s-IrO2(110). (e) Oot-H2Oot species and (f) a HOot/HOot pair resulting from molecular vs. dissociative chemisorption of an H2O molecule next to an Oot atom on IrO2(110). (g) HOot-H2Oot species and (h) HOot/HObr+ HOot species resulting from molecular vs. dissociative chemisorption of an H2O molecule next to a HOot group on IrO2(110).

Figure 4. Images showing computed structures of molecular and dissociated H2O chemisorbed on IrO2(110) and binding energies defined relative to the specified initial states. (a) H2Oot species and (b) a HObr/HOot pair resulting from molecular vs. dissociative chemisorption of an H2O molecule on s-IrO2(110). (c) HObr/HOot-H2Oot dimer and (d) HObr/HOot + HObr/HOot dimer with binding energies per H2O defined relative to 2H2O(g) + s-IrO2(110). (e) Oot-H2Oot species and (f) a HOot/HOot pair resulting from molecular vs. dissociative chemisorption of an H2O molecule next to an Oot atom on IrO2(110). (g) HOot-H2Oot species and (h) HOot/HObr+ HOot species resulting from molecular vs. dissociative chemisorption of an H2O molecule next to a HOot group on IrO2(110).

Computational results

Oxygen adsorption and dissociation on IrO2(110)

We performed DFT calculations to investigate the adsorption of O2 on s-IrO2(110). We examined O2 on atop and bridge sites of the Ircus row with different orientations and our calculations predict that the most favourable configuration of adsorbed O2 is a flat-lying, peroxy-species in which each O-atom bonds with an adjacent Ircus atom. The peroxy-species is the preferred configuration of molecularly-adsorbed O2, and has a computed binding energy of 182 kJ/mol at an O2 coverage of 0.125 ML (O-atom basis) in the 4 × 2 IrO2(110) model (see Computational Details), which corresponds to 0.50 ML of O-atoms along an Ircus row. The computed binding energy of the peroxy species is higher than the binding energies (~50 kJ/mol) suggested by the O2 TPD peaks observed below 200 K, due to both coverage effects as well as overbinding of molecularly adsorbed O2 predicted by the PBE functional [Citation29]. Notably, recent results provide evidence that PBE is more accurate for predicting the binding energies of adsorbed atomic oxygen [Citation30]. We find that the adsorbed peroxy species can dissociate to produce Oot atoms by overcoming a barrier of only 14 kJ/mol and that the exothermicity of reaction is 190 kJ/mol relative to the energy of a gas-phase O2 molecule. The calculations thus predict that O2 should dissociate efficiently at low coverage to generate Oot atoms. Figure S5 shows the computed energy pathway and molecular images for O2 dissociation on s-IrO2(110). We also find that the reaction exothermicity decreases to 157 kJ/mol for dissociation of 0.50 ML of adsorbed O2, demonstrating that the Oot-Ircus binding becomes weaker at high Oot coverage. These results are consistent with our experimental observations that O2 dissociates efficiently on IrO2(110) at low coverage and that the Oot-atoms become less stable with increasing coverage. We leave to future work a more detailed DFT study of O2 dissociation with increasing coverage, and instead focus on the effect of Oot atoms on H2 oxidation processes on the IrO2(110) surface.

Water chemisorption on IrO2(110)

We investigated H2O chemisorption on IrO2(110) using DFT to determine whether H2O adsorbs molecularly or dissociatively at low coverage and characterize the interactions between H2O and co-adsorbed Oot atoms and HOot groups. A key aim of these calculations is to obtain information that can aid in understanding the mechanism for H2 oxidation on O-rich IrO2(110). shows images of computed configurations of H2O and HO groups adsorbed next to different adsorbates on IrO2(110) and the corresponding binding energies determined using PBE. We also examined other configurations of different orientations for H2O and OH; however, we found that configurations forming hydrogen bonds (see ) are more energetically stable than other configurations.

Our calculations predict that an isolated H2Oot molecule binds on IrO2(110) by positioning its O-atom on-top of an Ircus atom and adopting an orientation that allows one of the H-atoms to hydrogen-bond with a neighbouring Obr atom (). The large binding energy of the chemisorbed H2O molecule (136 kJ/mol) is indicative of a strong, localized bonding interaction between H2O and IrO2(110). We predict that the dissociative chemisorption of H2O on s-IrO2(110) occurs by proton transfer to an Obr atom to produce a HObr/HOot pair in which the HOot group aligns its H-atom towards the Obr atom of the HObr group (). We find that dissociation of an isolated H2Oot molecule on s-IrO2(110) occurs with a negligible barrier, and is exothermic by 167 kJ/mol relative to the gaseous H2O reference state. According to DFT, the dissociative chemisorption of H2O on s-IrO2(110) is favorable over molecular adsorption by 31 kJ/mol at low coverage. We also investigated the stabilities of H2O-derived dimers on s-IrO2(110), and find that the dissociated dimer, 2(HObr/HOot), is more stable than the mixed HObr/HOot-H2Oot dimer on IrO2(110) (,)). However, the dissociated dimer is only ~3 kJ/mol more stable than the mixed dimer (per H2O molecule), compared with the larger stability of a single HObr/HOot pair relative to a single H2Oot species (~31 kJ/mol). This comparison reveals that hydrogen-bonding between HOot and H2Oot acts to stabilize the HObr/HOot-H2Oot dimer, and causes the mixed and dissociated dimers to have similar stabilities on IrO2(110).

Our calculations predict that an Oot atom stabilizes a neighbouring H2Oot molecule on IrO2(110) and that dissociative chemisorption of H2O is favoured over molecular adsorption on the O-rich surface. The H2Oot molecule adsorbed next to an Oot atom adopts a flat-lying orientation with one H-atom directed towards the Oot atom and the other directed towards an Obr atom, and has a binding energy of 178 kJ/mol relative to the gaseous H2O reference state (). Our results indicate that the Oot-H2Oot interaction stabilizes the H2Oot molecule by 42 kJ/mol relative to H2Oot adsorbed on s-IrO2(110) at low coverage. Dissociative chemisorption of H2O on the Oot-IrO2(110) surface produces a HOot/HOot pair and is exothermic by 197 kJ/mol (), corresponding to a stabilization of 19 kJ/mol over molecular adsorption. Our calculations predict that dissociative chemisorption is slightly favoured (by ~4 kJ/mol) over molecular adsorption when H2O adsorbs next to a HOot group on IrO2(110) (,)). This situation is similar to dimer formation on s-IrO2(110) (,)). Interestingly, we find that H2O adsorption next to a HOot group is more stabilizing than adsorption next to an HOot/HObr pair (,)), with the added stability equal to about 12 kJ/mol. Key findings of our calculations are that H2O dissociation is favoured on IrO2(110) at low H2O coverage, and that H2O achieves stronger binding on O-rich IrO2(110).

Comparison with prior studies reveals similar characteristics in H2O adsorption on IrO2(110) and RuO2(110). For example, the dissociative adsorption of a single H2O molecule is favoured over molecular adsorption on both oxides though DFT predicts that the HObr/HOot pair is only ~3 kJ/mol more stable than a single H2Oot molecule on RuO2(110) [Citation31] compared with 31 kJ/mol for IrO2(110). A recent study also shows that formation of a HOot/HOot pair is favoured during H2O adsorption on O-rich RuO2(110) with an exothermicity that is ~20 kJ/mol higher than adsorption of a H2O molecule that is separated from the Oot atom [Citation32]; the exothermicity is 60 kJ/mol for the analogous comparison of HOot/HOot vs. H2Oot on O-rich IrO2(110) (,)). The larger energy differences between molecular and dissociative H2O adsorption demonstrate that IrO2(110) is more active than RuO2(110) in promoting H2O deprotonation.

Mechanism for H2 oxidation on s-IrO2(110)

We recently investigated the oxidation of H2 on s-IrO2(110) using TPRS and DFT, and find that the reaction mechanism can be summarized using a few key steps as shown in . Our prior results demonstrate that H2 adsorbs strongly on Ircus sites (step 1) and that the molecularly-adsorbed H2 species (H2,ot) readily dissociates via H-transfer to an Obr atom, resulting in Hot and HObr species (step 2). We find that H2 binds to the Ircus atom through a strong dative bonding interaction, resulting in a molecularly-adsorbed H2 σ-complex with a computed binding energy of 78 kJ/mol. DFT calculations predict that the H2,ot species dissociates with a barrier of only 13 kJ/mol and that the reaction is exothermic by 113 kJ/mol. These energetics suggest that H2 dissociation on s-IrO2(110) occurs with high probability at temperatures as low as 85 K. According to DFT, Hot atom hopping to an Obr site is the predominant reaction step after H2 dissociation, and will nominally cause all the Hot atoms to convert to HObr groups prior to H2O formation. The hopping of a Hot atom to an Obr site is exothermic by 36 kJ/mol and features a barrier of 85 kJ/mol.

Table 1. Key reaction steps involved in H2 oxidation on s-IrO2(110) as identified by DFT. Energy barriers (Ea) and energy changes of reaction (ΔE) are shown for each step.

We predict that the recombination of HObr groups to produce gaseous H2O can occur by several pathways that are endothermic by between ~220 and 260 kJ/mol [Citation20]. These pathways are the most energetically demanding among the sequence of H2 oxidation steps that we identified, and are expected to control the overall rate of H2 oxidation to H2O over s-IrO2(110). The large barriers for HObr conversion to gaseous H2O are generally attributable to the preference for H2O dissociation and strong binding of HObr groups on s-IrO2(110) as well as energetic penalties associated with the creation of bridging oxygen vacancies (Ov). Our prior DFT calculations also predict that the barriers for H-atom diffusion along the Ircus and Obr rows are quite large, with values of 199 and 213 kJ/mol, respectively, and comparable to the barriers for H2O formation. An implication is that H-atoms will be effectively trapped near the original H2 dissociation site (Ircus and adjacent Obr atoms) at moderate temperature and that diffusion away from these sites can only begin to occur at appreciable rates at temperatures at which H2O formation and desorption also occurs at significant rates.

Mechanism for H2 oxidation on O-rich IrO2(110)

We used DFT to identify viable reaction steps that govern H2 oxidation on O-rich IrO2(110), and determine how Oot-atoms promote the conversion of adsorbed H2 to water. Our calculations reveal two major differences in the H2 oxidation mechanism on O-rich vs. stoichiometric IrO2(110). First, we find that the recombination of HO groups to produce gaseous H2O can occur by lower energy pathways in the presence of Oot atoms compared with s-IrO2(110). Our calculations also predict that the diffusion of H-atoms is highly facile from Obr to Oot atoms and also along the Oot rows. The fast diffusion favours the formation of HOot groups and thus water production by low energy pathways. Below, we describe the reaction steps involved in H2 oxidation on O-rich IrO2(110) as determined from DFT calculations.

lists key reaction steps involved in H2 oxidation on O-rich IrO2(110) as well as the energy barriers and heats of reaction determined from DFT calculations. Figures S6–S11 show energy diagrams and molecular images for each of these reaction steps. We focused mainly on systems with an H2 coverage of 0.25 ML and an Oot coverage of 0.25 or 0.50 ML along an Ircus row. Exploring numerous H and Oot coverages and configurations lies outside of the scope of the present study. Our calculations predict similar energetics for the first three reaction steps on O-rich and stoichiometric IrO2(110). Adsorption of H2 on an Ircus atom located next to an Oot atom produces a strongly-bound H2 σ-complex with a binding energy of 85 kJ/mol. The calculations predict that dissociation of the H2,ot species is highly facile via reaction with either an Obr or an Oot atom, with barriers less than 20 kJ/mol in each case, but that dissociation via H-transfer to an Oot atom is more exothermic by 30 kJ/mol. The calculations also predict that Hot hopping onto a Oot atom to produce an HOot group must overcome a barrier of 89 kJ/mol and is exothermic by 84 kJ/mol. This reaction barrier is similar to that required for Hot to hop onto an Obr atom (85 kJ/mol) but the reaction exothermicity is higher for the Hot + Oot reaction compared with the Hot + Obr reaction (89 vs. 36 kJ/mol). Overall, we find similar barriers but larger exothermicity for HOot vs. HObr formation as a result of H2 dissociation and Hot atom hopping.

Table 2. Key reaction steps involved in H2 oxidation on O-rich IrO2(110) as identified by DFT. Energy barriers (Ea) and energy changes of reaction (ΔE) are shown for each step.

Our calculations predict that OH recombination reactions are the most energetically-demanding reaction steps during H2 oxidation on O-rich IrO2(110), demonstrating that the H2O formation steps are rate-limiting in H2 oxidation on O-rich as well as stoichiometric IrO2(110). Compared with s-IrO2(110), the H2O formation reactions are more facile on O-rich IrO2(110) and occur by the reverse pathways identified computationally for the dissociative chemisorption of H2O on IrO2(110) (). This finding is consistent with our conclusion that H2O evolution is desorption-limited during H2 oxidation on O-rich IrO2(110). Among the surface configurations that we studied, the recombination of an HOot group with an HObr group to generate gaseous H2O and an Obr atom is the most favoured pathway for H2O formation on O-rich IrO2(110) and has a thermochemical barrier of 167 kJ/mol (, step 4A). This reaction is the reverse of the dissociative chemisorption of an H2O molecule on s-IrO2(110) (). also shows that the recombination of HOot groups to generate gaseous H2O and an Oot atom has an overall thermochemical barrier of 196 kJ/mol (step 4B), and may proceed through the formation of a H2Oot-Oot intermediate. We speculate that HOot + HObr recombination contributes mainly at low Oot coverage, but that recombination of adjacent HOot groups becomes increasingly important at higher Oot coverage. Such a change in the dominant H2O formation pathway may explain, at least in part, the upshift in the desorption-limited H2O TPRS feature that we observe with increasing initial Oot coverage ().

Our calculations predict that H-atom hopping is highly facile from Obr to Oot atoms and also between adjacent Oot atoms. Hydrogen atom hopping between Oot atoms has a computed barrier of 17 kJ/mol, and H-atom hopping from an Obr to an Oot atom has a barrier of 9 kJ/mol and is exothermic by 27 kJ/mol (). These fast diffusion processes can enable a quasi-equilibrium to be established between the HOot and HObr groups in local regions of the surface. Equilibrium between HOot and HObr groups favors formation of the HOot species at the temperatures of interest (< 750 K) because the HObr + Oot → Obr + HOot reaction is significantly exothermic (ΔE = −27 kJ/mol). The extent of such equilibration may depend sensitively on the local configurations and species coverages, since other surface species diffuse more slowly compared with H-atom hopping from Obr/Oot to Oot atoms [Citation20].

Our DFT calculations suggest a scenario wherein H2 dissociates efficiently on O-rich IrO2(110) to produce Hot and HO species. The resulting surface H-atoms will redistribute via site-hopping processes to produce a mixture of HOot and HObr groups prior to H2O formation, where the equilibrium composition of this mixture favours HOot groups below ~750 K. Water formation then ensues by several HO recombination steps, the most favourable of which is H-transfer from an HObr to a HOot group and evolution of H2O (step 4A, ). The contributions from the various HO recombination reactions will depend on the relative H and Oot coverages. For saturation H2 coverage, as investigated in our experiments, an increasing fraction of H2O will be produced by the more facile HOot + HObr reaction as the Oot coverage initially increases, while HObr recombination will make a decreasing contribution. However, since HOot groups are energetically favoured over HObr groups, water formation by HOot + HOot recombination (step 4B, ) will begin to dominate above an intermediate Oot coverage. This scenario would cause the H2O production rate at a given temperature to initially increase with increasing Oot coverage as the HOot + HObr reaction makes an increasing contribution, but the H2O production rate would then decrease with further increases in Oot coverage because the barrier for HOot + HOot recombination is higher than that for HOot + HObr recombination (). Such behaviour is qualitatively consistent with our experimental observations that the desorption-limited H2O TPRS peak replaces the broad reaction-limited H2O feature and also shifts towards higher temperature with increasing Oot coverage ().

The present study reveals that H2 oxidation involves analogous steps on O-rich IrO2(110) and RuO2(110). Prior work specifically demonstrates that H-transfer from Obr to Oot atoms is facile and leads to the production of H2Oot species at low temperature, followed by the desorption-limited evolution of H2O during TPRS [Citation12,Citation15,Citation17]. A recent study of H2O adsorption on O-rich RuO2(110) supports interpretations of the H2 oxidation mechanism on RuO2(110) as well as IrO2(110), demonstrating that H-transfer from Obr to Oot atoms involves a low barrier (~20 kJ/mol) and that HOot groups are favoured over HObr groups by ~16 kJ/mol [Citation32]. Those authors also predict that a HOot/HOot pair is more stable than a HObr/HOot pair on RuO2(110) by ~17 kJ/mol. These findings are similar to the results of the present study, though our calculations predict larger energy differences between HOot and HObr groups on IrO2(110). Notably, Jacobi et al. also find that the desorption-limited H2O TPRS peak upshifts from ~400 to 425 K with increasing Oot coverage during H2 oxidation on RuO2(110) [Citation12]. We find that the desorption-limited H2O TPRS peak shifts from 490 to 550 K with increasing Oot coverage on IrO2(110) (). Comparison suggests a higher energy barrier for H2O desorption as well as larger stabilization of an HOot/HOot vs. an HObr/HOot pair on IrO2(110) compared with RuO2(110). These findings are consistent with DFT results, and support the interpretation that desorption-limited evolution of H2O from O-rich RuO2(110) also occurs by the mechanism that we propose here, namely, that HObr/HOot recombination is important during H2 oxidation at low Oot coverage, but that the higher-barrier HOot/HOot recombination step becomes increasingly dominant as the Oot coverage increases.

Summary

We investigated the adsorption and oxidation of H2 on O-rich IrO2(110) using TPRS and DFT calculations. We find that the dissociative chemisorption of hydrogen occurs efficiently on O-rich IrO2(110) at low temperature and that Oot atoms facilitate the subsequent oxidation of adsorbed hydrogen. Our results show that Oot atoms promote the desorption-limited evolution of H2O during H2 oxidation on IrO2(110) while suppressing reaction-limited H2O production arising from recombination of HObr groups. The desorption-limited H2O TPRS peak initially appears at 490 K but shifts to 550 K with increasing Oot coverage, demonstrating that Oot atoms stabilize adsorbed water species. We find that the H2 uptake on IrO2(110) decreases with increasing Oot coverage and interpret this behaviour as evidence that H2 dissociation initiates from the adsorbed H2 σ-complex state and thus requires Ircus sites to proceed under the conditions studied.

Our DFT calculations predict that Oot atoms promote H2 oxidation on IrO2(110) and that H2O evolution via desorption-limited processes is rate-determining in H2 oxidation on O-rich IrO2(110). The calculations predict that dissociative chemisorption of H2O is favoured over molecular adsorption on both stoichiometric and O-rich IrO2(110), and produces HObr/HOot and HOot/HOot pairs, respectively. Desorption-limited evolution of H2O from IrO2(110) thus occurs by the recombination of HObr/HOot or HOot/HOot pairs, with the latter involving a higher energy barrier (196 vs. 167 kJ/mol) due to the greater stability of HOot compared with HObr groups. According to DFT, H2 adsorbs molecularly on Ircus sites as a strongly-bound σ-complex and dissociates efficiently at low temperature to produce Hot and HO groups. Our calculations suggest that the H-atoms rapidly redistribute on the surface to generate a mixture of HObr and HOot groups, with equilibrium favouring HOot formation at temperatures below at least 750 K. Water evolution subsequently occurs on O-rich IrO2(110) by the desorption-limited recombination of HObr/HOot and HOot/HOot pairs. Our results suggest that the lower energy HObr/HOot recombination pathway will be important at low Oot coverages, while the higher-energy HOot/HOot recombination pathway will become dominant at higher Oot coverages due the preferable formation of HOot groups prior to water formation. This prediction is consistent with our experimental finding that the desorption-limited H2O TPRS peak shifts to higher temperature with increasing Oot coverage.

Supporting information

Estimates of adsorbate coverage and product yields from TPRS; H2O TPD spectra obtained from O2-exposed IrO2(110) at 90 K; H2O TPD spectra from s-IrO2(110) after H2O adsorption at 85 K; TPRS for H2 oxidation on IrO2(110) with Oot-layer prepared at 300 K; H2 oxidation yields as a function of initial Oot coverage prepared by heating to selected temperatures vs. O2 adsorption at 300 K; Computed energy pathway and molecular images for O2 dissociation on IrO2(110); Computed energy diagrams and molecular images for key steps involved in H2 oxidation on O-rich IrO2(110).

Supplemental material

Supplemental Material

Download PDF (888.1 KB)

Acknowledgments

We acknowledge the Ohio Supercomputing Center for providing computational resources. We gratefully acknowledge financial support for this work provided by the Department of Energy, Office of Basic Energy Sciences, Catalysis Science Division through Grant DE-FG02-03ER15478.

Disclosure statement

No potential conflict of interest was reported by the authors.

Supplementary material

Supplemental data can be accessed here.

Additional information

Funding

This work was supported by the Office of Basic Energy Sciences [DE-FG02-03ER15478].

References

  • Smith RDL, Sporinova B, Fagan RD, et al. Facile photochemical preparation of amorphous iridium oxide films for water oxidation catalysis. Chem Mater. 2014;26(4):1654–1659.
  • Slavcheva E, Radev I, Bliznakov S, et al. Sputtered iridium oxide films as electrocatalysts for water splitting via PEM electrolysis. Electrochim Acta. 2007;52(12):3889–3894.
  • Rasten E, Hagen G, Tunold R. Electrocatalysis in water electrolysis with solid polymer electrolyte. Electrochim Acta. 2003;48(25–26):3945–3952.
  • Boodts JCF, Trasatti S. Hydrogen evolution on iridium oxide cathodes. J Appl Electrochem. 1989;19(2):255–262.
  • Liang Z, Li T, Kim M, et al. Low-temperature activation of methane on the IrO2(110) surface. Science. 2017;356(6335):298–301.
  • Bian YX, Kim M, Li T, et al. Facile dehydrogenation of ethane on the IrO2(110) surface. J Am Chem Soc. 2018;140(7):2665–2672.
  • Martin R, Kim M, Franklin A, et al. Adsorption and oxidation of propane and cyclopropane on IrO2(110). Phys Chem Chem Phys. 2018;20:29264–29273.
  • Weaver JF. Surface chemistry of late transition metal oxides. Chem Rev. 2013;113:4164−4215.
  • Over H. Surface chemistry of ruthenium dioxide in heterogeneous catalysis and electrocatalysis: from fundamental to applied research. Chem Rev. 2012;112(6):3356–3426.
  • Kim YD, Seitsonen AP, Wendt S, et al. Characterization of various oxygen species on an oxide surface: ruO2(110). J Phys Chem B. 2001;105(18):3752–3758.
  • Wang JH, Fan CY, Sun Q, et al. Surface coordination chemistry: dihydrogen versus hydride complexes on RuO2(110). Angew Chem Int Ed. 2003;42(19):2151–2154.
  • Jacobi K, Wang Y, Ertl G. Interaction of hydrogen with RuO2(110) surfaces: activity differences between various oxygen species. J Phys Chem B. 2006;110(12):6115–6122.
  • Knapp M, Crihan D, Seitsonen AP, et al. Hydrogen transfer reaction on the surface of an oxide catalyst. J Am Chem Soc. 2005;127(10):3236–3237.
  • Knapp M, Crihan D, Seitsonen AP, et al. Unusual process of water formation on RuO2(110) by hydrogen exposure at room temperature. J Phys Chem B. 2006;110(29):14007–14010.
  • Knapp M, Crihan D, Seitsonen AP, et al. Complex interaction of hydrogen with the RuO2(110) surface. J Phys Chem C. 2007;111(14):5363–5373.
  • Wei YY, Martinez U, Lammich L, et al. Formation of metastable, heterolytic H-pairs on the RuO2(110) surface. Surf Sci. 2014;619:L1–L5.
  • Wei YY, Martinez U, Lammich L, et al. Atomic-scale view on the RuO2 formation reaction from H2 on O-Rich RuO2(110). J Phys Chem C. 2014;118(48):27989–27997.
  • Henderson MA, Dahal A, Dohnalek Z, et al. Strong Temperature dependence in the reactivity of H2 on RuO2(110). J Phys Chem Lett. 2016;7(15):2967–2970.
  • Henderson MA, Mu RT, Dahal A, et al. Light makes a surface banana-bond split: photodesorption of molecular hydrogen from RuO2(110). J Am Chem Soc. 2016;138(28):8714–8717.
  • Li T, Kim M, Liang Z, et al. Dissociative chemisorption and oxidation of H2 on the stoichiometric IrO2(110) surface. Top Catal. 2018;61(5–6):397–411.
  • Blochl PE. Projector augmented-wave method. Phys Rev B. 1994;50(24):17953–17979.
  • Kresse G, Hafner J. Abinitio Hellmann-Feynman molecular-dynamics for liquid-metals. J Non-Cryst Solids. 1993;156:956–960.
  • Kresse G. Ab-initio molecular-dynamics for liquid-metals. J Non-Cryst Solids. 1995;193:222–229.
  • Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made simple. Phys Rev Lett. 1996;77(18):3865–3868.
  • Henkelman G, Uberuaga BP, Jonsson H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J Chem Phys. 2000;113(22):9901–9904.
  • Pogodin S, Lopez N. A more accurate kinetic monte carlo approach to a monodimensional surface reaction: the interaction of oxygen with the RuO2(110) surface. ACS Catal. 2014;4(7):2328–2332.
  • Wang HY, Schneider WF. Effects of coverage on the structures, energetics, and electronics of oxygen adsorption on RuO2(110). J Chem Phys. 2007;127(6):064706.
  • Li T, Kim M, Rai R, et al. Adsorption of alkanes on stoichiometric and oxygen-rich RuO2(110). Phys Chem Chem Phys. 2016;18(32):22647–22660.
  • Pan L, Weaver JF, Asthagiri A. First principles study of molecular O2 adsorption on the PdO(101) surface. Top Catal. 2017;60(6–7):401–412.
  • Kim M, Pan L, Weaver JF, et al. Initial reduction of the PdO(101) surface: role of oxygen vacancy formation kinetics. J Phys Chem C. 2018;122(45):26007–26017.
  • Mu RT, Cantu DC, Glezakou VA, et al. Deprotonated water dimers: the building blocks of segmented water chains on rutile RuO2(110). J Phys Chem C. 2015;119(41):23552–23558.
  • Nguyen MT, Mu RT, Cantu DC, et al. Dynamics, stability, and adsorption states of water on oxidized RuO2(110). J Phys Chem C. 2017;121(34):18505–18515.