Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Thermal synthesis of electron deficient oxygen species on crystalline IrO2

E. A. Carbonio *ab, F. Sulzmann b, D. Teschner bc, J. J. Velasco-Vélez bcd, M. Hävecker ac, A. Knop Gericke bc, R. Schlögl abc and T. Jones *be
aCatalysts for Energy, Energy Materials In-situ Laboratory (EMIL), Helmholtz-Zentrum Berlin für Materialien und Energie GmbH, BESSY II, Albert-Einstein-Straße 15, 12489 Berlin, Germany. E-mail: carbonio@fhi-berlin.mpg.de
bDepartment of Inorganic Chemistry, Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4-6, 14195 Berlin, Germany. E-mail: tejones@lanl.gov
cDepartment of Heterogeneous Reactions, Max Planck Institute for Chemical Energy Conversion, Stiftstraße 34-36, Mülheim an der Ruhr 45470, Germany
dALBA Synchrotron Light Source, Cerdanyola del Vallés (Barcelona) 08290, Spain
eTheoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA

Received 24th July 2023 , Accepted 26th November 2023

First published on 11th December 2023


Abstract

Water splitting is a promising technology in the path towards complete renewable energy within the hydrogen economy but overcoming the sluggishness of the oxygen evolution reaction (OER) is a major challenge. Iridium-based oxides remain the most attractive materials for the OER under acidic conditions since they offer the combination of activity and stability. Gaining knowledge about how these materials have such an ability is of great interest to develop improved electrocatalysts for the OER. Among the different iridium-based oxides the materials with high concentrations of electron deficient oxygen (OI−) have been shown to have higher OER activity, however, they also have high dissolution rates, seemingly due to the presence or formation of IrIII species. In contrast, rutile-type IrO2, which does not contain IrIII species, has high dissolution resistance but the OER activity remains comparatively low as only low coverages of OI− species are formed under OER. The apparent link between OI− and IrIII species that comes from these observations has yet to be proven. In this work, using ab initio thermodynamics and in situ X-ray photoelectron and absorption spectroscopy we show that the same electrophilic OI− species that appear on Ir-based oxides under OER can be formed on IrIV+δ by mild thermal oxidation of rutile-type IrO2, without the presence IrIII species.


Introduction

The oxygen evolution reaction (OER) is a kinetic bottleneck in water electrolysis, requiring high-performance catalysts to achieve low overpotentials. This problem is especially acute for the anodes of proton exchange membrane (PEM) based electrolyzers. An apparent inverse relationship between stability and activity,1,2 results in only Ir-based materials providing the required combination of stability and activity.3,4 This has made Ir-based materials the subject of intense scrutiny as highlighted in a number recent review.5–9 Electrochemical studies of Ir-based catalysts have shown that, while the thermodynamically stable rutile-type IrO2 shows the higher dissolution resistance,2,10 hydrated X-ray amorphous oxyhydroxides (IrOx) show improved activity with respect to their crystalline counterparts at the expense of stability. For instance, a unique oxygen ligand environment facilitates water oxidation in hole-doped IrNiOx core–shell electrocatalysts.11–14 The improved activity of amorphous IrOx was attributed to the presence of electron deficient oxygen, or OI−, based on studies combining X-ray absorption spectroscopy (XAS) at the O K-edge combined with density functional theory calculations (DFT,)12,15–17 where the OI− species gives a characteristic resonance at ca. 529 eV in the O K-edge spectrum. The reduced stability of amorphous materials may be attributable to the presence of IrIII species.10,18,19 It is then important to understand if there is a link between OI− and IrIII. Unfortunately, the X-ray amorphous nature of IrOx makes further structural characterization of the observed OI− difficult in XAS studies. And while TEM investigations of amorphous materials have helped to identify key structural features related to activity,20,21 there is still a pressing need to understand whether there is a difference in active species on amorphous and rutile-type Ir-oxides in order to improve catalyst performance while minimizing stability losses.

The original assignment of OI− in IrOx to undercoordinated μ2-O species (O bound to two rather than three Ir-atoms, see Fig. S1)12,15,17 has proven to be successful in describing a range of operando XAS results at both the O K-edge and Ir L2,3-edges.10,11,13,14,18,22–26 This species does not require the presence of IrIII, and could therefore be formed without the deleterious stability effects associated with reduced Ir species. If this assignment is correct, μ2-O should form on rutile-type IrO2 under gas-phase conditions; μ2-O is expected to be stable on rutile-type IrO2 under vacuum conditions,27–30 though, to the best of our knowledge, it has not been spectroscopically observed at room temperature or on crystalline powder catalysts.12,15,17,29,30 Here we address this apparent discrepancy in the assignment of OI− and its possible relationship with IrIII through in situ X-ray photoelectron spectroscopy, in situ O K-edge XAS, and chemical testing of well-calcined rutile-type IrO2 combined with DFT calculations. In agreement with its assignment to μ2-O, we find that the electron deficient-O species can be formed on rutile-type IrO2 by mild-oxidation (conditions), albeit with a coverage that is less than that observed for IrOx.

Methodology

Sample preparation

A crystalline rutile IrO2 (Sigma Aldrich, 99.9%) powder, known to have little OER activity,15 was calcined at 1073 K for 50 hours under 1 bar O2 to ensure rutile-type phase purity.10,16 For some experiments (specified in the figures), an amorphous IrOx (Alfa Aesar) powder was calcined at 1123 K to ensure crystallization and rutile-type phase purity.10,16 Characterization of both types of powders before and after calcination has been previously published by our group.15,16 Unless stated otherwise, (calcined) IrO2 Sigma Aldrich (99.9%) has been used. For the near ambient pressure (NAP) XPS/XAS measurements, 5 mm diameter pellets where prepared by pressing 50 mg of material under 2 tons for 2 minutes. The samples were placed in a sapphire sample holder between two stainless-steel plates (the front facing plate contained an opening for measurements). Heating was done with an IR laser on the backside stainless-steel plate that was directly in contact with the sample. The sample temperature was measured with a K-type thermocouple placed on top of the sample pellet and held between the sample and the front stainless-steel plate. The temperature was controlled by adjusting the laser power using a PID feedback loop.

For NAP XPS/XAS measurements the samples were kept in UHV, then treated in 0.25 mbar O2 at 573 K and finally exposed to 0.25 mbar CO. For the CO treatment, the sample was first cooled to 473 K under 0.25 mbar O2 and then cooled to 313 K under UHV. The sample was then kept in the load lock (base pressure 10−7 mbar). The main chamber was filled with 0.25 mbar CO, and after stabilization of the pressure, the load lock valve was open to expose the sample to the 0.25 mbar CO. This methodology has been shown to be a reliable way of testing CO oxidation in the NAP-XPS setup.17 For all experiments 99.9999% O2 and 99.97% CO from Westfalen were used.

X-ray absorption and X-ray photoelectron spectroscopy

Measurements were done at the UE56/2-PGM beamline at the BESSY II synchrotron radiation light source in Berlin, Germany (in low alpha mode). The details of the beamline can be found elsewhere.31

For the XP difference spectrum, a Shirley background with a 10% offset on the high binding energy side was subtracted from the spectra, which were normalized afterwards. The Ir 4f core-level was measured with a photon energy of 150 eV, which is ca. 85–90 eV kinetic energy (KE). This gives an estimated inelastic mean free path (IMPF) of approximately 0.4 nm;32 note the lattice spacing in the (110) direction is 0.3 nm,33 making the 150 eV photons surface sensitive. The XA spectra were measured by Auger electron yield (AEY) using electrons at a fixed KE of 514 eV (and pass energy of 50 eV). The probing depth of this measurement is approximately 2–3 nm,17 or ca. 6–10 atomic layers in the (110) direction. XPS and XAS were measured using a Phoibos 150 hemispherical analyzer from SPECS GmbH.

On-line quadrupole mass spectrometry

A quadrupole mass spectrometer (QMS) from Pfeiffer (directly connected to the main reaction XPS/XAS chamber through a leak valve) was used on-line as a means of detecting CO2 produced from CO oxidation. Ions with m/z = 44, 28, 32, 18 were recorded to follow CO2, CO, O2 and H2O, respectively.

Computational details

Spin polarized density functional theory calculations were performed using the Quantum ESPRESSO package.34,35 All calculations employed the PBE exchange and correlation potential36 using PAW datasets from the PSLibrary37 with a kinetic energy (charge density) cutoff of 60 Ry (600 Ry). Rutile-type IrO2 surfaces were modeled using 8-layer symmetric slabs with (110) surface termination with the central 4-layer fixed during geometry optimizations. The (110) surface was chosen as it is the lowest energy surface and has been observed to be the dominant facet for rutile IrO2 subjected to high temperature treatments.38–40 Since this surface has only in an oxidation state above III, to explore lower oxidation state Ir, we have included surfaces with oxygen vacancies together with bulk heterogenite-type IrOOH and corundum-type Ir2O3 (Fig. S7). Other facets and the effect of Ir vacancies have been previously studied by our group.10,15,17,18 These do not affect the outcome of this publication. For the surfaces, periodic images were separated by ca. 15 Å of vacuum. A k-point mesh equivalent to (6 × 3) for the (1 × 1) surface unit cell was employed with cold smearing using a smearing parameter of 0.01 Ry.41 For bulk Ir2O3 and IrOOH, (6 × 6 × 2) and (8 × 8 × 8) k-point meshes were employed, respectively. O K-edge spectra were computed using the single particle approach implemented in the XSpectra package.42 Following previous publications,15 O K-edge spectra were computed in the absence of a core hole. Core level binding energies were computed by the ΔSCF method43 using (2 × 4) surfaces, and (2 × 2 × 2) supercells for the bulk structures. These were aligned to experiment using a ΔSCF calculation on a (4 × 4 × 4) supercell of bulk rutile-type IrO2, where the ΔSCF shift was aligned to experiment following the procedure outlined in a previous publication.44 All computed spectra of the surface species were then aligned using this bulk IrO2 reference. The Ir 4f spectra were computed using a Hopfield model wherein the final state contained an Ir 4f corehole.18

The surface phase diagrams were computed by way of ab initio atomistic thermodynamics.45,46 To avoid the well-known over binding of O2,gas predicted by the PBE exchange and correlation potential, the oxygen reference energy was taken from bulk rutile-type IrO2 as: EO2 = EIrO2EIr − ΔHf,IrO2, where EO2 is the total energy of O2, EIrO2 is the computed total energy bulk rutile-type IrO2; EIr is the computed total energy of bulk fcc Ir; ΔHf,IrO2 is the experimental heat of formation of the rutile-type oxide.47 The enthalpic and entropic corrections for the gas-phase oxygen chemical potential were taken from the NIST JANAF tables,48 allowing us to express the oxygen chemical potential as: μO2(T, pO2) = EO2 + [small mu, Greek, tilde]O2(T, p0) + kBT[thin space (1/6-em)]ln(pO2/p0), where p0 is standard pressure. Here, [small mu, Greek, tilde]O2(T, p0) is the tabulated enthalpic and entropic contributions for gas-phase O2 at standard pressure. The final term accounts for pressure variations of the ideal-gas reservoir. We assume water will be the major hydrogen source in air or under NAP-XPS conditions, which allows us to write the hydrogen chemical potential as: μH2(T, pO2, pH2O) = μH2O(T, pH2O) − 1/2μO2(T, pO2), where μH2O(T, pH2O) = EH2O + [small mu, Greek, tilde]H2O(T, p0) + kBT[thin space (1/6-em)]ln(pH2O/p0).43 In this case, the energy of the gas-phase water molecule, EH2O, was computed at the Γ-point in (20 × 20 × 20) Å3 box and the enthalpic and entropic contributions, [small mu, Greek, tilde]H2O(T, p0), were again taken from the NIST JANAF tables. This choice of μH2 can be seen to be conservative as compared to hydrocarbon sources that may be present in air or under typical NAP-XPS conditions, since the employed definition involves the heat of formation of water. The predicted stability of OH structures is then expected to be a lower bound for what is found experimentally. In all cases the entropic contributions from the adsorbates have been ignored. From previous work we anticipate this will lead to an underestimation of the thermal stability of adsorbed species by about 100–150 K,44 which will not change the results presented herein.

Results and discussion

Highly electrochemically active amorphous IrOx contains electron deficient oxygen species, or OI−, assigned to doubly coordinated oxygen atoms (μ2-O) (see structure in Fig. S1) that have not been observed on rutile-type IrO2 powder under UHV15–17 and are present only in small amounts under OER conditions.10 It is expected that such species could form under gas-phase oxygen. To test the conditions under which this will occur, we turn to ab initio thermodynamics. We find that at room temperature rutile-type IrO2 is predicted to be covered by μ1-OH/μ2-OH species (see structures in Fig. S1) under 0.25 mbar O2,gas even when only trace (10−8 mbar) H2Ogas is present (Fig. 1), however, μ2-O is predicted to form at higher temperatures under these partial pressures due to the entropic cost of hydrogen adsorption. And critically, for the (110) surface terminated IrO2, these μ2-O form in the absence of IrIII. This predicted presence of μ2-O in the absence of IrIII can be tested experimentally by way of NAP-XPS since these are realizably characterized by NAP-XPS.
image file: d3cy01026k-f1.tif
Fig. 1 Calculated surface phase diagram for the surface species on rutile IrO2 (110) under 0.25 mbar O2 and 10−8 mbar H2O, where the zero of energy is the bare (110) surface.

This finding is consistent with previous publications on IrO2 thin layers grown on single crystals that have shown that surface species were observed to mainly consist of protonated O (–OH) species, that form rapidly due to background H2 content, unless the thin film is kept at higher temperatures.30 Here, unlike the μ2-OH sites, the μ1-OH sites are not expected to be oxidized to μ1-O under mbar gas-phase conditions when trace water (or other H sources) is present (Fig. 1). Under experimentally accessible conditions at mbar O2 pressures (Fig. 1), the μ1-O phase is only metastable (below ca. 700 K) owing to the strong O–H bond of μ1-OH. Furthermore, the (hydro)peroxo species argued to participate in OER and postulated as possibly present in IrOx (ref. 49 and 50) appear at higher energies and are unstable at temperatures above ca. 350 K under gas-phase conditions with mbar pressures and should therefore not be observed. Thus, NAP-XPS offers an ideal means of testing the assignment of OI− and its relationship with IrIII. We investigated this experimentally by means of NAP-XPS and XAS.

The Ir 4f spectrum of the IrO2 pellet (Fig. 2a) shows the typical line shape and binding energy (BE) for rutile-type IrO2.15,16,29,30 The O-K edge spectrum of rutile-type IrO2 (Fig. 2b) shows a sharp resonance at ca. 530 eV, characteristic of the lattice μ3-O.16,17 Computed Ir 4f and O-K edge spectra for the different species are shown in Fig. 2c and d (and S2). As OH species have a very low resonance intensity per atom at the O K-edge,10,23 the μ1-OH and μ2-OH species — which are predicted to cover the surface and have resonances at ca. 529.5 and 530.5 eV (Fig. S2) — will have very low intensity compared to the major bulk contribution of μ3-O. A clear resonance from μ1-OH and μ2-OH species is then not expected, nor is it seen. No resonance at ca. 529 eV (characteristic of μ2-O species on iridium oxides) is observed at room temperature under UHV (see also Fig. S3), in agreement with our ab initio thermodynamics predictions (Fig. 1) and previous findings.16


image file: d3cy01026k-f2.tif
Fig. 2 (a) Ir 4f of rutile IrO2 in UHV at RT, under 0.25 mbar O2 at 573 K, and difference spectrum. (b) AEY O K-edge of rutile IrO2 in UHV at RT and under 0.25 mbar O2 at 573 K. Vertical lines indicate the corresponding species giving rise to each resonance. (c) Computed Ir 4f spectra for bulk-IrIV in rutile-type IrO2, surface Ir–μ2-OH and surface Ir–μ2-O. Inset shows a zoom in to the Ir 4f7/2. (d) Computed O K-edges for μ3-O (bulk-O), μ2-OH and μ2-O species.

At elevated temperatures the surface phase of the rutile-type oxide is predicted to transform from μ2-OH to μ2-O (Fig. 1). From the computed spectra, this phase transition is predicted to introduce a higher binding energy component and additional asymmetry in the Ir 4f spectrum (Fig. 2c), and a ca. 529 eV resonance in the O K-edge spectrum (Fig. 2d and S2). As predicted, the Ir 4f XP spectrum of the rutile-type IrO2 shows a slight broadening upon heating to 573 K under 0.25 mbar O2. This broadening can be assigned to the IrIV+δ formed upon dehydrogenation of the μ2-OH sites to form an electron deficient O-species, O(I+δ)−, which also yields IrIV+δ due to the covalency of the Ir–O bond (in agreement with earlier observations).10,23,30 For simplicity, we call this electron deficient O-species OI− throughout this manuscript (as named in the literature).12,13,15–17,50 The change becomes more apparent when we analyze the difference spectrum (Fig. 2a), which resembles computed spectra for Ir bound to μ2-O (Fig. 2c). At the O-K edge, a resonance at ca. 529 eV can also be seen to develop when heating the sample under O2 (Fig. 2b, see also S4 and S5). The species formed by mild thermal oxidation has the same spectroscopic characteristics as the species found in amorphous IrOx (under UHV) and iridium oxyhydr(oxides) under OER.10,15–17,23

The surface phase transition from μ1-OH/μ2-OH to μ1-OH/μ2-O is also predicted to show changes in the projected density of states (pDOS) on the rutile-type IrO2 (Fig. 3). The O 2p pDOS (Fig. 3a) shows an increased intensity at ca. 0.2, 2 and 3 eV below the Fermi energy for the surface with μ2-O species. At the same time, Ir shows a loss of 5d states at about 1.3 eV below the Fermi energy for the Ir–μ2-O species (Fig. 3b) due to its oxidation. The computed pDOS can be compared to valence band (VB) measurements of the rutile-type IrO2. These have been taken under UHV and under 0.25 mbar O2 at 573 K, as shown in Fig. 3c and d, respectively. The VB were measured at two different photon energies, 150 eV and 420 eV. The first one (150 eV) is near the Cooper minimum of iridium, thus maximizing the sensitivity to oxygen species due to the comparatively high O 2p cross-section; the O 2p cross-section is ca. 3 times that of the Ir 5d at this energy.51 For the VB measured at 420 eV the Ir cross section is ca. 5 times larger than that of the O 2p, and the spectra are dominated by the Ir 5d contribution.51 The VB measured at 150 eV under oxygen at 573 K (Fig. 3c) shows an increased intensity between 0–4 eV when compared to the VB measured under UHV at RT. At the same time, the VB measured at 420 eV (Fig. 3d) under oxygen at 573 K shows a slight loss of intensity at ca. 1.5 eV when compared to the VB measured under UHV at RT. These changes although small, are reproducible (see Fig. S6) and consistent with the predicted changes in the pDOS for the phase transition from μ2-OH to μ2-O, and agree well with the XPS and XAS results (Fig. 2). Computed O K-edge spectra, pDOS and the calculated BE for bulk IrIII compounds and rutile IrO2 surfaces containing IrIII and IrIII+δ (for both μ1-OH/μ2-OH and μ1-OH/μ2-O surfaces) can be found in the ESI (see Fig. S7–S9, and Table S1). All these species have a BE equal or lower than the IrIV bulk component for the Ir 4f, and most present little or no t2g (ca. 529 eV) resonance in the O K-edge. The pDOS shows that for μ2-O formation on IrIII a loss in the O 2p intensity is expected (Fig. S9), opposite to the observed gain in intensity observed experimentally (Fig. 3c and d) and predicted for μ2-O on IrIV+δ (Fig. 3a and b). The combined observed changes in Ir 4f, O K-edge and VB for the rutile type IrO2 under O2 are entirely consistent solely with the computed Ir 4f and O K-edge spectra and pDOS for μ2-O on IrIV+δ, thus supporting the assignment of the formed species to μ2-O formed in the absence of IrIII.


image file: d3cy01026k-f3.tif
Fig. 3 DOS for O 2p (a) and Ir 5d (b) states for different O surface species on rutile-type IrO2. VB measured at 150 eV (c) and 420 eV (d) for rutile-type IrO2 in UHV and RT and under 0.25 mbar O2 at 573 K.

As noted above, the computed surface phase diagram (Fig. 1) predicts μ1-OH remains stable across the entire range of oxygen chemical potentials accessible under NAP-XPS conditions and the experiments show no sign of the 528 eV resonance associated with the formation of μ1-O (ref. 10 and 23) (see Fig. S2) argued to be active in electrochemical water oxidation. Moreover, the VB measurements also show no sign of the formation of μ1-O (see Fig. S10). Thus, we can then expect that μ2-O can be formed by thermal oxidation in addition to electrochemical oxidation, whereas μ1-O formation requires more oxidizing conditions than those accessible in this work, in good agreement with the computed surface phase diagram.

To verify the μ2-O formed on rutile-type IrO2 in the absence of IrIII has the chemical behavior associated with OI− we can also explore its reactivity with CO. The μ2-O species linked to the high OER activity on amorphous IrOx have been shown to have an electrophilic nature and to be active in room temperature CO oxidation,17,50 without any pre-treatment. Moreover, iridium oxyhydroxides with high concentration of μ2-O species were shown to lose OER activity after partial titration of the μ2-O species with CO, showing μ2-O plays a prominent role in the OER performances of iridium based (oxyhydr)oxides,50 In contrast, rutile-type IrO2 without any pre-treatment is not active in CO oxidation at room temperature.17 With this in mind, we turn to testing the chemistry of the thermally produced species on rutile-type IrO2.

After producing μ2-O on rutile-type IrO2 through O2 treatment at 573 K, the sample was cooled to room temperature and placed under vacuum. It was then exposed to 0.25 mbar CO at room temperature and the O K-edge was monitored. Upon CO exposure CO2 was produced (Fig. 4a). At the same time, the resonance developed at ca. 529 eV at the O-K edge for the O2-treated rutile-type IrO2 can be seen to disappear, indicating the loss of μ2-O (Fig. 4b and S11). The Ir 4f spectrum can be seen to lose the previously gained high-energy broadening (Fig. 4c) and a similar difference-spectrum (from that in Fig. 2a) is obtained when comparing the rutile-type IrO2 treated under O2 and the same sample after exposure to CO (Fig. 4c). A difference spectrum between the Ir 4f spectra of the rutile-type IrO2 before the O2 treatment and after exposure to CO confirms that the spectrum of the rutile-type IrO2 after CO exposure is the same as before the O2 treatment (Fig. 4d). That is, the μ2-O formed on IrIV+δ by the thermal oxygen treatment was titrated by CO. Moreover, the VB (Fig. 4e) measured at 150 eV (sensitive to the O 2p) shows the loss of intensity related to the μ2-O species (compare Fig. 4e to 3c), and the VB (Fig. 4f) measured at 420 eV (dominated by the Ir 5d) shows a slight increase of intensity related to the loss of IrIV+δ (compare Fig. 4f to 3b). The changes are seen to be again (although small) reproducible (Fig. S12 and S13). With all the above we confirm that the species produced on rutile-type IrO2 by (mild) thermal oxidation has the same chemical nature as the OI− present on/in highly active amorphous iridium oxides. Thus, we have successfully produced electron deficient (electrophilic) oxygen species on rutile-type IrO2 by thermal oxidation of μ2-OH species (accompanied by oxidation of IrIV to IrIV+δ) without the need of IrIII species. Depth profiling (Fig. S14) reveals that the species produced under the O2 treatment are located in the (near) surface of the IrO2. This is in agreement with μ2-OH deprotonation on rutile-type IrO2 under OER occurring mainly on the surface,10 rendering only low concentrations of OI− on the less active rutile-type IrO2.


image file: d3cy01026k-f4.tif
Fig. 4 Here: (a) CO2 signal (m/z = 44) from on-line QMS for rutile-type IrO2 exposed to 0.25 mbar of CO at RT, before and after O2 treatment at 573 K. (b) O K-edge of rutile-type IrO2 under 0.25 mbar O2 at 573 K and in UHV after CO at RT. (c and d) Ir 4f of rutile-type IrO2 under 0.25 mbar O2 at 573 K, in UHV after CO at RT and in UHV as loaded. The difference spectra are also shown. (e and f) VB of rutile-type IrO2 measured at 150 eV (e) and 420 eV (f) under 0.25 mbar O2 at 573 K and in UHV after CO at RT.

To address why μ2-O species have not been spectroscopically observed on rutile-type IrO2 at room temperature previously we return to ab initio thermodynamics. Even at O2 partial pressure as high as ca. 210 mbar and H2O partial pressure as low as 10−12 mbar, μ2-OH is the thermodynamically preferred species at room temperature (Fig. S15). This means that when rutile-type IrO2 is exposed to any water traces, or other hydrogen sources, μ2-OH will form at room temperature. In our experiments we found that during the cooling ramp after the O2 treatment of the rutile-type IrO2, the resonance at 529 eV (which corresponds to μ2-O) decreased with time (even under O2 which contains a maximum of 0.5 ppm H2O, i.e. on the order of 10−7 mbar H2O in the 0.25 mbar O2). With the chosen protocol (see Methodology section) the μ2-O species remained on the surface long enough to probe their chemistry by exposing the sample to CO, as we have shown above.

This observation and the ab initio atomistic thermodynamics results agree with the strong dissociative adsorption of H2O (ref. 27 and 52) and the previously observed fast protonation of oxygen surface species under UHV by background H2 (ref. 30) reported on IrO2 surfaces. Thus, any rutile-type IrO2 in the presence of small (≥0.5 ppm) amounts of water will have a hydroxylated surface with protonated μ2-OH and μ1-OH species. And while we have developed a protocol to prepare μ2-O via thermal oxidation, the stronger O–H bond of μ1-OH makes preparation of μ1-O more challenging and we were unable to produce this species under gas-phase NAP-XPS conditions. Experiments as described here verify the reactive oxyl species is only formed under higher oxygen chemical potential.

Under these controlled preparation conditions, it becomes possible to address an unresolved aspect of electron deficient oxygen or OI−. The electron deficient μ2-O species has been described in hydrated amorphous IrOx to be formed in a matrix of mixed IrIII/IV oxide, though no direct link between IrIII and OI− was established.15,16 And although IrIII might be present under OER conditions, and a wide variety of defects and iridium oxidation states present in amorphous materials may facilitate lattice oxidation and Ir dissolution,53–55 here we have shown that under O2 at 573 K the electron deficient μ2-O species appearing on crystalline IrO2 are formed on IrIV+δ without the need for IrIII. It thus appears that the IrIII present in amorphous IrOx is unrelated to OI−, which has important consequences for Ir-based OER catalysts as IrIII is unstable.19

Our work shows increasing the OI− coverage through doping11 or variation in local atomic structure20,21 need not introduce IrIII. This may then offer a route to higher-performance stable OER catalysts. Here, designing complex oxides with stable basic-coordination sites acting as proton-acceptors next to the reactive Ir species could be a good approach to better OER catalysts.

Conclusions

Using ab initio atomistic thermodynamics, we have predicted the surface phase diagram of rutile-type IrO2 under gas phase O2 and trace amounts of water. We computed the spectroscopic properties of these surface phases and confirmed the ab initio atomistic thermodynamics predictions by means of XPS and XAS. We found mild thermal oxidation of rutile-IrO2 is sufficient to form an electron deficient O-species, μ2-O bound to IrIV+δ sites. This species is not present at room temperature over rutile-type IrO2 due to the high hydrogen affinity of μ2-O, which becomes hydroxylated even at extremely low (10−12 mbar) water (and high O2) partial pressure. With CO titration experiments we confirmed the chemical nature of the produced species, allowing us to show μ2-O has the same spectroscopic and chemical properties associated with the OI− species whose coverage is related to OER activity of Ir-based materials. By preparing this OI− under controlled conditions, however, we were able to show its presence does not require IrIII, which could offer a route to higher-performance stable OER catalysts.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We thank the Helmholtz-Zentrum Berlin für Materialien und Energie GmbH BESSYII synchrotron light source for providing support of the in situ electron spectroscopy activities of the Fritz Haber Institute der Max Planck Gesellschaft, and we thanks the Max-Planck Gesellschaft for generous founding. TJ acknowledges support from the Laboratory Directed Research and Development program of Los Alamos National Laboratory under project number 20240061 DR. Open Access funding provided by the Max Planck Society.

References

  1. N. Danilovic, R. Subbaraman, K. C. Chang, S. H. Chang, Y. J. J. Kang, J. Snyder, A. P. Paulikas, D. Strmcnik, Y. T. Kim, D. Myers, V. R. Stamenkovic and N. M. Markovic, J. Phys. Chem. Lett., 2014, 5, 2474–2478 CrossRef CAS.
  2. S. Cherevko, S. Geiger, O. Kasian, N. Kulyk, J. P. Grote, A. Savan, B. R. Shrestha, S. Merzlikin, B. Breitbach, A. Ludwig and K. J. J. Mayrhofer, Catal. Today, 2016, 262, 170–180 CrossRef CAS.
  3. M. Carmo, D. L. Fritz, J. Merge and D. Stolten, Int. J. Hydrogen Energy, 2013, 38, 4901–4934 CrossRef CAS.
  4. M. Schalenbach, G. Tjarks, M. Carmo, W. Lueke, M. Mueller and D. Stolten, J. Electrochem. Soc., 2016, 163, F3197–F3208 CrossRef CAS.
  5. T. Naito, T. Shinagawa, T. Nishimoto and K. Takanabe, Inorg. Chem. Front., 2021, 8, 2900–2917 RSC.
  6. Z. J. Chen, X. G. Duan, W. Wei, S. B. Wang and B. J. Ni, Nano Energy, 2020, 78, 105270 CrossRef CAS.
  7. C. Wang, F. F. Lan, Z. F. He, X. F. Xie, Y. H. Zhao, H. Hou, L. Guo, V. Murugadoss, H. Liu, Q. Shao, Q. Gao, T. Ding, R. B. Wei and Z. H. Guo, ChemSusChem, 2019, 12, 1576–1590 CrossRef CAS PubMed.
  8. H. Jang and J. Lee, J. Energy Chem., 2020, 46, 152–172 CrossRef.
  9. H. Wu, Y. Wang, Z. Shi, X. Wang, J. Yang, M. Xiao, J. Ge, W. Xing and C. Liu, J. Mater. Chem. A, 2022, 10, 13170–13189 RSC.
  10. R. V. Mom, L. J. Falling, O. Kasian, G. Algara-Siller, D. Teschner, R. H. Crabtree, A. Knop-Gericke, K. J. J. Mayrhofer, J.-J. Velasco-Vélez and T. E. Jones, ACS Catal., 2022, 12, 5174–5184 CrossRef CAS.
  11. H. N. Nong, T. Reier, H. S. Oh, M. Gliech, P. Paciok, T. H. T. Vu, D. Teschner, M. Heggen, V. Petkov, R. Schlogl, T. Jones and P. Strasser, Nat. Catal., 2018, 1, 841–851 CrossRef CAS.
  12. V. Pfeifer, T. E. Jones, J. J. Velasco Vélez, R. Arrigo, S. Piccinin, M. Hävecker, A. Knop-Gericke and R. Schlögl, Chem. Sci., 2017, 8, 2143–2149 RSC.
  13. V. A. Saveleva, L. Wang, D. Teschner, T. Jones, A. S. Gago, K. A. Friedrich, S. Zafeiratos, R. Schlogl and E. R. Savinova, J. Phys. Chem. Lett., 2018, 9, 3154–3160 CrossRef CAS.
  14. H. N. Nong, L. J. Falling, A. Bergmann, M. Klingenhof, H. P. Tran, C. Spori, R. Mom, J. Timoshenko, G. Zichittella, A. Knop-Gericke, S. Piccinin, J. Perez-Ramirez, B. R. Cuenya, R. Schlogl, P. Strasser, D. Teschner and T. E. Jones, Nature, 2021, 589, E8 CrossRef CAS PubMed.
  15. V. Pfeifer, T. E. Jones, J. J. V. Velez, C. Massue, M. T. Greiner, R. Arrigo, D. Teschner, F. Girgsdies, M. Scherzer, J. Allan, M. Hashagen, G. Weinberg, S. Piccinin, M. Havecker, A. Knop-Gericke and R. Schlogl, Phys. Chem. Chem. Phys., 2016, 18, 2292–2296 RSC.
  16. V. Pfeifer, T. E. Jones, J. J. V. Velez, C. Massue, R. Arrigo, D. Teschner, F. Girgsdies, M. Scherzer, M. T. Greiner, J. Allan, M. Hashagen, G. Weinberg, S. Piccinin, M. Havecker, A. Knop-Gericke and R. Schlogl, Surf. Interface Anal., 2016, 48, 261–273 CrossRef.
  17. V. Pfeifer, T. E. Jones, S. Wrabetz, C. Massue, J. J. V. Velez, R. Arrigo, M. Scherzer, S. Piccinin, M. Havecker, A. Knop-Gericke and R. Schlogl, Chem. Sci., 2016, 7, 6791–6795 RSC.
  18. J. J. Velasco-Velez, E. A. Carbonio, C. H. Chuang, C. J. Hsu, J. F. Lee, R. Arrigo, M. Havecker, R. Z. Wang, M. Plodinec, F. R. Wang, A. Centeno, A. Zurutuza, L. J. Falling, R. V. Mom, S. Hofmann, R. Schlogl, A. Knop-Gericke and T. E. Jones, J. Am. Chem. Soc., 2021, 143, 12524–12534 CrossRef.
  19. O. Kasian, J. P. Grote, S. Geiger, S. Cherevko and K. J. J. Mayrhofer, Angew. Chem., Int. Ed., 2018, 57, 2488–2491 CrossRef.
  20. E. Willinger, C. Massue, R. Schlogl and M. G. Willinger, J. Am. Chem. Soc., 2017, 139, 12093–12101 CrossRef CAS PubMed.
  21. M. Elmaalouf, M. Odziomek, S. Duran, M. Gayrard, M. Bahri, C. Tard, A. Zitolo, B. Lassalle-Kaiser, J. Y. Piquemal, O. Ersen, C. Boissiere, C. Sanchez, M. Giraud, M. Faustini and J. Peron, Nat. Commun., 2021, 12, 3935 CrossRef CAS.
  22. J. J. Velasco-Velez, T. E. Jones, V. Streibel, M. Havecker, C. H. Chuang, L. Frevel, M. Plodinec, A. Centeno, A. Zurutuza, R. Wang, R. Arrigo, R. Mom, S. Hofmann, R. Schlogl and A. Knop-Gericke, Surf. Sci., 2019, 681, 1–8 CrossRef CAS.
  23. L. J. Frevel, R. Mom, J.-J. Velasco-Vélez, M. Plodinec, A. Knop-Gericke, R. Schlögl and T. E. Jones, J. Phys. Chem. C, 2019, 123, 9146–9152 CrossRef CAS.
  24. T. Reier, Z. Pawolek, S. Cherevko, M. Bruns, T. Jones, D. Teschner, S. Selve, A. Bergmann, H. N. Nong, R. Schlogl, K. J. J. Mayrhofer and P. Strasser, J. Am. Chem. Soc., 2015, 137, 13031–13040 CrossRef CAS PubMed.
  25. P. Strasser, Acc. Chem. Res., 2016, 49, 2658–2668 CrossRef CAS PubMed.
  26. H. N. Nong, H. P. Tran, C. Spöri, M. Klingenhof, L. Frevel, T. E. Jones, T. Cottre, B. Kaiser, W. Jaegermann, R. Schlögl, D. Teschner and P. Strasser, Z. Phys. Chem., 2020, 234, 787–812 CrossRef CAS.
  27. D. Gonzalez, J. Heras-Domingo, S. Pantaleone, A. Rimola, L. Rodriguez-Santiago, X. Solans-Monfort and M. Sodupe, ACS Omega, 2019, 4, 2989–2999 CrossRef CAS.
  28. T. Li, M. Kim, Z. Liang, A. Asthagiri and J. F. Weaver, Top. Catal., 2018, 61, 397–411 CrossRef CAS.
  29. M. J. S. Abb, T. Weber, D. Langsdorf, V. Koller, S. M. Gericke, S. Pfaff, M. Busch, J. Zetterberg, A. Preobrajenski, H. Grönbeck, E. Lundgren and H. Over, J. Phys. Chem. C, 2020, 124, 15324–15336 CrossRef CAS.
  30. R. Martin, M. Kim, C. J. Lee, V. Mehar, S. Albertin, U. Hejral, L. R. Merte, E. Lundgren, A. Asthagiri and J. F. Weaver, J. Phys. Chem. Lett., 2020, 11, 7184–7189 CrossRef CAS PubMed.
  31. K. J. S. Sawhney, F. Senf, M. Scheer, F. Schäfers, J. Bahrdt, A. Gaupp and W. Gudat, Nucl. Instrum. Methods Phys. Res., Sect. A, 1997, 390, 395–402 CrossRef CAS.
  32. S. Tanuma, C. J. Powell and D. R. Penn, Surf. Interface Anal., 1994, 21, 165–176 CrossRef CAS.
  33. IrO2 Crystal Structure: Datasheet from “PAULING FILE Multinaries Edition – 2012” in SpringerMaterials (https://materials.springer.com/isp/crystallographic/docs/sd_0457154); Springer-Verlag Berlin Heidelberg & Material Phases Data System (MPDS), Switzerland & National Institute for Materials Science (NIMS), Japan.
  34. P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. Dal Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari and R. M. Wentzcovitch, J. Phys.: Condens. Matter, 2009, 21, 395502 CrossRef PubMed.
  35. P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. B. Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, M. Cococcioni, N. Colonna, I. Carnimeo, A. Dal Corso, S. de Gironcoli, P. Delugas, R. A. DiStasio, A. Ferretti, A. Floris, G. Fratesi, G. Fugallo, R. Gebauer, U. Gerstmann, F. Giustino, T. Gorni, J. Jia, M. Kawamura, H. Y. Ko, A. Kokalj, E. Kucukbenli, M. Lazzeri, M. Marsili, N. Marzari, F. Mauri, N. L. Nguyen, H. V. Nguyen, A. Otero-de-la-Roza, L. Paulatto, S. Ponce, D. Rocca, R. Sabatini, B. Santra, M. Schlipf, A. P. Seitsonen, A. Smogunov, I. Timrov, T. Thonhauser, P. Umari, N. Vast, X. Wu and S. Baroni, J. Phys.: Condens. Matter, 2017, 29, 465901 CrossRef CAS.
  36. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  37. A. Dal Corso, Comput. Mater. Sci., 2014, 95, 337–350 CrossRef CAS.
  38. G. Novell-Leruth, G. Carchini and N. López, J. Chem. Phys., 2013, 138, 194706 CrossRef.
  39. J. Rossmeisl, Z. W. Qu, H. Zhu, G. J. Kroes and J. K. Norskov, J. Electroanal. Chem., 2007, 607, 83–89 CrossRef CAS.
  40. T. D. Nguyen, G. G. Scherer and Z. C. J. Xu, Electrocatalysis, 2016, 7, 420–427 CrossRef CAS.
  41. N. Marzari, D. Vanderbilt, A. De Vita and M. C. Payne, Phys. Rev. Lett., 1999, 82, 3296–3299 CrossRef CAS.
  42. C. Gougoussis, M. Calandra, A. P. Seitsonen and F. Mauri, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80, 075102 CrossRef.
  43. E. Pehlke and M. Scheffler, Phys. Rev. Lett., 1993, 71, 2338–2341 CrossRef CAS PubMed.
  44. T. E. Jones, T. C. R. Rocha, A. Knop-Gericke, C. Stampfl, R. Schlogl and S. Piccinin, Phys. Chem. Chem. Phys., 2015, 17, 9288–9312 RSC.
  45. K. Reuter and M. Scheffler, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 65, 035406 CrossRef.
  46. C. F. S. Codeco, A. Y. Klyushin, E. A. Carbonio, A. Knop-Gericke, R. Schlogl, T. Jones and T. C. R. Rocha, Phys. Chem. Chem. Phys., 2022, 24, 8832–8838 RSC.
  47. E. H. P. Cordfunke, Thermochim. Acta, 1981, 50, 177–185 CrossRef CAS.
  48. M. W. Chase Jr., NIST-JANAF thermochemical tables, J. Phys. Chem. Ref. Data, Monogr., 1998 DOI:10.18434/T42S31.
  49. J. J. V. Velez, D. Bernsmeier, T. E. Jones, P. Zeller, E. Carbonio, C. H. Chuang, L. J. Falling, V. Streibel, R. V. Mom, A. Hammud, M. Havecker, R. Arrigo, E. Stotz, T. Lunkenbein, A. Knop-Gericke, R. Krahnert and R. Schlogl, Faraday Discuss., 2022, 236, 103–125 RSC.
  50. C. Massue, V. Pfeifer, M. van Gastel, J. Noack, G. Algara-Siller, S. Cap and R. Schlogl, ChemSusChem, 2017, 10, 4786–4798 CrossRef PubMed.
  51. J. J. Yeh and I. Lindau, At. Data Nucl. Data Tables, 1985, 32, 1–155 CrossRef.
  52. D. González, M. Sodupe, L. Rodríguez-Santiago and X. Solans-Monfort, Nanoscale, 2021, 13, 14480–14489 RSC.
  53. A. Loncar, D. Escalera-López, S. Cherevko and N. Hodnik, Angew. Chem., Int. Ed., 2022, 61, e202114437 CrossRef.
  54. A. Zagalskaya and V. Alexandrov, J. Phys. Chem. Lett., 2020, 11, 2695–2700 CrossRef CAS PubMed.
  55. L. A. She, G. Q. Zhao, T. Y. Ma, J. Chen, W. P. Sun and H. G. Pan, Adv. Funct. Mater., 2022, 32, 2108465 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cy01026k

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.