Ir local structure and its influence on proton penetration for IrO x and IrO2.
In this study, hydrous, amorphous iridium oxide films were prepared by electrodeposition on FTO substrates, following a well-established procedure.27, 31 (see Supplementary Method). Rutile IrO2 films were obtained by thermal annealing of the electrodeposited IrOx films at 550 ℃ for 8 hours, as confirmed by XRD (Supplementary Fig. 1C). Scanning electron microscopy (SEM) images show that both amorphous and rutile IrO2 films consist of 100–200 nm size nanoparticles (Supplementary Fig. 1A and 1B). However, in the as-deposited state, IrOx shows a combination of Ir3+ and Ir4+ redox states, while the IrO2 is dominated by Ir4+, as determined by X-ray photoelectron spectroscopy (XPS) (Supplementary Fig. 1D), consistent with previous work 32, 33. This is further supported by ex-situ X-ray absorption near-edge structure (XANES) measurements at the Ir L3 edge. The average oxidation state is determined by the position of the white line edge, which primarily corresponds to transition from occupied 2p to empty 5d states.12 As shown in Supplementary Fig. 2, the d-band holes for amorphous IrOx and rutile IrO2 are around 4.4 and 5.0, corresponding to average oxidation states of around 3.4 and 4.0, respectively (see Supplementary Methods for XANES measurements and analysis). The fitted extended X-ray absorption fine structure (EXAFS) results (Fig. 1C and Supplementary Fig. 3, Table 1) show that both samples have a characteristic Ir–O interatomic distance of ~ 2.0 Å, with comparable first shell coordination number of 5.8 (± 0.5) and 6.0 (± 0.7) for amorphous IrOx and rutile IrO2, respectively. However, rutile IrO2 shows a stronger signal from Ir-Ir1 and Ir-Ir2 (between 3 and 4 Å interatomic distance) while these signals are very weak in amorphous IrOx. Thus, these data suggest short range ordered [IrO6] units are dominant for amorphous IrOx whilst a longer range ordered structure is present in rutile IrO2, as schematically shown in Fig. 1A and 1B.
Cyclic voltammograms of rutile IrOx show one order of magnitude higher redox current densities, when normalised to geometric area than IrO2. IrOx also shows higher activity than IrO2 as reflected in the overpotential required to achieve 0.5 mA/cmgeom2 (~ 210 mV and 330 mV for IrOx and IrO2 respectively, Fig. 1D). IrOx shows two distinct broad redox peaks at around 0.8 V and 1.2 V vs RHE. In contrast, no distinct redox peak can be observed for IrO2. In order to determine whether the increased current density on amorphous IrOx is only a consequence of higher permeability of electrolyte within the porous structure (resulting in a higher density of catalytically active sites) or is also a result of higher proton penetration through the bulk of the oxide (resulting in participation of subsurface iridium sites), a combination of TOF-SIMS and deuterium isotope labeling was used. The film was impregnated and covered with paraffin wax, using an adapted method from Hadden and co-workers.34 This method eliminates the effects of varying porosity on the ion incorporation (See Fig. 1E and Supplementary Figs. 4 and 5 for details) - and thus probes only the effect of surface exchange and bulk diffusion directly. Samples were cycled in 0.1 M DClO4 (99.8% D) from 0.66 V to 1.5 V. The depth of proton penetration during water oxidation was assessed by the difference in deuterium signal intensity of the cycled samples and the samples exposed to electrolyte but without electrochemical cycling. Figure 1F shows that amorphous IrOx cycled in DClO4 has a higher D/H ratio than the sample without cycling, for the first 100 seconds of sputtering, corresponding to a distance of 20 nm. This signal then drops to the same level after 100 s (see Supplementary Figs. 6 and 7 and Table 2 for details). In contrast, no obvious difference in D/H ratio can be observed before and after cycling for rutile IrO2. These results indicate that protons penetrate far deeper in amorphous IrOx than rutile IrO2 under CV cycling, thus offering more iridium sites for redox reactions and possibly for water oxidation. We conjecture that the short range ordered structure in amorphous IrOx is more flexible and hence allows more proton intercalation than the long range ordered – and hence presumably more rigid – structure in rutile IrO2. This is also consistent with the observation from atom probe tomography by Mayrhofer, Cherevko, Gault, Kasian and co-workers that indicate that hydrous IrOx undergoes a greater degree of oxygen exchange with the electrolyte during OER than rutile IrO2 where exchanges is limited to within 2.5 nm of the surface.35, 36
Identification And Quantification Of Redox Transitions On Amorphous And Rutile Iridium Oxides
Next, we probe the redox chemistry of both materials using time-resolved operando UV-Vis spectroscopy and stepped potential spectroelectrochemistry (see Supplementary method for details). On increasing the potential, amorphous IrOx shows broad absorption bands at 600 nm, 800 nm and 500 nm in potential ranges of ~ 0.66 V to 1.0 V vs RHE, ~ 1.1 V to 1.32 V vs RHE and > 1.32 V vs RHE respectively (Fig. 2a). Three distinct differential absorption spectral shapes can be obtained by a subtraction of adjacent spectra (Supplementary Fig. 9), and are denoted as redox 1, redox 2 and redox 3, in accordance with our previous work.27, 28 The differential absorption of the redox transitions in IrO2 are very similar to those for IrOx, with only a slight shift of the dominant absorption band for redox 1 to 700 nm (Supplementary Fig. 10). However, the intensity of the optical absorption is 6 times lower in the case of IrO2 (Fig. 2b), consistent with this more crystalline and rigid structure having a lower density of redox active Ir centres, as discussed above. These results show that although IrOx and IrO2 have significantly different CV curves, they undergo the same redox processes, with both generating three similar redox states within the applied potential range.
To analyse the contribution of each redox species to the observed spectra as a function of potential, we deconvoluted the absorption spectra through a linear combination fitting process, which gives the absorption intensity from each individual redox transition as a function of potential (see Supplementary Figs. 11 and 12 for details). The obtained absorption intensities were then converted to an area density of redox species by applying a potential step and correlating the charge passed to the absorption intensity (see Supplementary Figs. 13 and 14 for details). The potential limits for the stepped potential spectroelectrochemistry measurements were selected to cover ranges in which the differential spectral shape is not changing (Supplementary Figs. 9 and 10), thereby allowing us to assume that only one redox transition takes place during the potential step. The redox peak positions obtained from these optical signals for the amorphous and crystalline iridium oxide were found to be at about 0.82 and 0.9 V for redox 1, 1.22 and 1.26 V for redox 2, and 1.47 and 1.57 V for redox 3, respectively (Supplementary Table 3, Fig. 2d and 2e). These optically determined redox transition waves match well with the less well resolved redox transitions observed in the linear sweep voltammetry curves for amorphous IrOx (grey line in Fig. 2d, top panel). The positions of redox 1 and 2 are also in agreement with the redox peak positions on amorphous IrOx and perovskite SrIrOx determined by Geiger et al (both are at around 0.8 and 1.2 V)10. Notably, the redox current for redox 3 corresponding to the optical data at OER relevant potentials is masked by the OER current measured electrochemically in the CV. A redox peak at similar potential (1.63 V) has also been observed by Suntivich and co-workers on a IrO2 (110) surface.30 It is clear that the amorphous IrOx undergoes the redox 3 process (circa 1.3 V) at approximately 100 mV lower potential than rutile IrO2 (circa 1.4 V, see Supplementary Fig. 15 for direct absorption signal comparison). This suggests the differences in coordination between Ir centres in amorphous and crystalline iridium oxides result in stronger binding of oxygenated intermediates in IrOx compared to IrO2.
The concentration of the redox transitions on amorphous IrOx was obtained by integration of the above redox waves; they show typical sigmoid shapes with increasing potential (bottom panel of Fig. 2d), with similar concentrations of redox active states participating in all three redox processes. These electroadsorption isotherms can be modelled with a Frumkin isotherm with R2 values as high as 0.99, as opposed to a simple Langmuir isotherm (dashed line in Fig. 2d, bottom panel, see supplementary Fig. 16–17 for details). These excellent fits indicate the existence of lateral interactions between the redox states, as assumed in the Frumkin model. Mathematically, in a Frumkin electroadsorption model, the redox transition free energy ∆G0redox is a function of the redox state coverage, i.e. ∆G0redox(θ)= ∆G0redox (θ=0) + rθ, where r (in eV) is the interaction energy of the absorbates at full coverage, θ is the coverage of the absorbates (0 < θ < 1). θ was calculated by assuming that saturating concentrations observed in our spectroelectrochemical analyses indicates full coverage, and r was obtained by fitting the θ vs U data (see supplementary Fig. 16–17 for details). The same analysis procedure was also applied to rutile IrO2. The saturated redox concentration in IrOx (~ 3.5x1016 cm− 2) is more than 6 times higher than that for IrO2 (~ 5.5x1015 cm− 2), which we attribute to the higher proton penetration in addition to the higher porosity within the structure (Fig. 2d). The fitted interaction energy r and the half-coverage potential U(θ=1/2) for both amorphous and rutile iridium oxides are labelled in Fig. 2D and 2E. Interestingly, we found that the adsorbate-adsorbate interaction energy r for redox 3 on amorphous IrOx and rutile IrO2 are similar, both at about 0.15 eV, indicating the interaction strength between the absorbates involved in this redox transition are similar despite the changes in the Ir coordination environment between amorphous and crystalline iridium oxides.
The UV-vis absorption changes in Fig. 2 can be assigned to changes in intervalence charge transfer within the Ir d orbitals, which correlates to the oxidation of Ir and the surface absorbates coordinated to the Ir centre.37–39 To verify the nature of the surface absorbates corresponding to these redox transitions, we use density functional theory (DFT) calculations (see Computational Methods in the supplementary information). We model the rutile IrO2 (110) surface, which consists of rows of Ir atoms coordinated by bridging oxygens separated by rows of coordinatively unsaturated (CUS) sites (inset in Fig. 2G, see details in supplementary Fig. 18). The adsorption of *H, *H2O, *OH and *O on the CUS site and *H on the bridging oxygen (*Hb) was considered at full coverage in order to determine the most stable surface adsorbate at different potentials (Supplementary Fig. 19). The interaction between the absorbates was explored by calculating the free energy of the stepwise transition between the most stable adsorbates in a unit cell with three CUS sites, i.e at different coverages of the species. For selected surfaces the OER was modelled, two different reaction pathways were considered as described in previous work40(Supplementary Fig. 20). Based on these energies, the redox transition potential at half coverage U(θ=1/2), and the interaction energy, r, were calculated and plotted as adsorption isotherms in Fig. 2G. The calculated energies at half coverage for *OH (1.05 VDFT−RHE) and *O (1.52 VDFT−RHE) on the CUS site of rutile IrO2 match well with experimental half-wave redox potentials for redox 2 (1.26 V) and redox 3 (1.57 V) (Supplementary Fig. 20, supplementary table 5). Our calculations are also in agreement with previous work which has attributed the redox transitions at around 1.2 V and 1.63 V on IrO2 (110) surfaces in 0.1 M HClO4 to *OH and *O adsorption on the CUS site, respectively.30 The surface transition for redox 1 was considered as *H desorbing from the CUS site or *H2O dissociating to form *OH on the CUS site, but the calculated potential values for both transitions (< 0.19 VDFT−RHE) do not match closely with experimental results of 0.90V, even if the possibility of an intermediate structure with some CUS sites covered by *H2O is considered (see supplementary Fig. 19 for details). It is therefore likely that our DFT calculations do not fully capture the nature of redox transition 1. Based on the calculations and the charge observed in the experiment, we assign the redox transitions to (also see schematic in Fig. 2C):
Redox 1: * H2O (cus) + *Hb → *OH(cus) + *Hb + H+ + e− (1)
Redox 2: *OH(cus) + *Hb→ *OH(cus) + H+ + e− (2)
Redox 3: *OH(cus) → *O(cus) + H+ + e− (3)
The surface of amorphous IrOx cannot be modeled by periodic DFT calculations. Instead, as a representative of a more open IrOx structure with larger separation between the adsorbates, the (001) surface of hollandite IrO2 is modelled and calculations are performed for the transitions (OH(cus) + *Hb → *OH(cus) + H+ + e− ) and (*OH(cus) →*O(cus) + H++ e−).41 The energy of the redox transitions 2 (0.97 V) and 3 (1.34 V) on hollandite IrOx are lower than rutile IrO2 (Supplementary Fig. 21), in agreement with our experimental observations.
Our DFT calculations also capture the interaction strength between *OH absorbates on both the rutile and the hollandite surfaces, showing that the calculated redox potential increase from 0.89 V at 1/3 coverage to 1.20 V at full coverage of *OH for the rutile structure, and from 0.80 to 1.13 V for the hollandite structure. (see details of the interaction parameter calculation in Supplementary Fig. 22–23 and supplementary table 5). However, the interaction of *O cannot be fully captured in this model. The calculated interaction energies of *O are only about 0.02 eV and 0.08 eV for rutile and hollandite IrO2, respectively, considerably lower than the experimentally observed value of 0.15 eV. This observation is in agreement with recent work suggesting that the repulsive interactions for *O species can only be captured by including solvent effects in the DFT model as the *O interactions might propagate through the water layer 26. The above experimental and theoretical results indicate the key role of repulsive interactions between the adsorbates, which give rise to a Frumkin electroadsorption isotherms (Fig. 2D, E), and change the free energy of elementary steps in OER as a function of coverage.
Determining The Intrinsic Water Oxidation Kinetics
Next, we investigated how the free energy of the elementary steps controls the reaction rate on amorphous IrOx and rutile IrO2. We used two methods to estimate the intrinsic rate of the rate-determining step as a function of potential (and thus also as a function of *O coverage). Firstly, the rate of O2 generation (obtained from the current) was divided by the concentration of redox state 3. This gives the active-state-normalized rate of O2 release, which is equivalent to the intrinsic rate of the RDS (denoted as TOF (O2 redox3− 1 s− 1)). Secondly, we measured the decay kinetics of the accumulated states when the potential was released from an oxygen-evolving potential to open circuit using transient time-resolved UV-Vis absorption spectroscopy.28, 42, 43 This experiment is shown in Fig. 3A, where two spectrally distinct decay phases were observed during open circuit decay – a fast decay component (initial 1 s) and a slow decay component (100–200 s) (see supplementary Fig. 24 for spectra of this two decay components). The differential absorption spectra for these two phases are in excellent agreement with the differential absorption spectra of redox 3 and redox 2, respectively (Supplementary 25), indicating the fast decay of absorption primarily arises from the decay of redox state 3 species (Fig. 3B), while slow decay mainly arises from the decay of redox state 2. Thus, this experiment can be interpreted as described in Fig. 3B: a site in the active state proceeds along the reaction path by nucleophilic attack from water molecular forming O-O bond followed by release of oxygen. Here, charge is balanced by other active states by accepting a proton and becoming reduced to regenerate *OH on the surface (since no charge is passed through the external circuit during open circuit). If the time constant of this active state decay (i.e.: fast phase) is τ, it implies an intrinsic rate of the RDS of 1/ 4τ of oxygen per second (see supplementary information for further discussion). By varying the applied potential and probing the decay when the cell is switched to open circuit, we are able to control the concentration of active states and investigate its effect on the intrinsic rate of the rate-determining step (denoted as TOF(1/4τ)) as a function of potential (detailed measurement of τ and the calculation of TOF(1/4τ) can be seen in Supplementary Figs. 26 and 27). Figure 3C shows the active state (*O) coverage and the intrinsic RDS rate of the RDS for amorphous IrOx and rutile IrO2. We find that while rutile IrO2 has fewer active states at a given potential, it has the same intrinsic rate of the RDS as measured by O2 evolution rate and slightly higher intrinsic rate of the RDS as measured by open-circuit decay.
Given that the coverage of the absorbates in the Frumkin isotherm plays a key role in the binding energetics of the adsorbates and thus controls the reactivity of the states, we next compare the TOF3 (O2 redox3− 1 s− 1) and TOF(1/4τ) for amorphous IrOx and rutile IrO2 as a function of the coverage of *O (Fig. 3D). We note that for both iridium oxides, log[TOF (1/4τ)] increases roughly linearly with the *O coverage in the observed coverage range. This result is similar to the observation in a recent work by Nong et al, where they found the log(current) linearly increase with the total charge and suggested from DFT that the activation energy of the RDS decreases linearly with the increase of *O coverage because of the long-range interaction among the *O species26. Here, by experimentally tracking the intrinsic rate of RDS as a function of *O coverage, we observe similar results. We also note that the slopes of log[TOF (1/4τ)] vs *O coverage in Fig. 3D are similar in amorphous IrOx and rutile IrO2, consistent with our observation that they have similar absorbate-absorbate interaction energies r. The repulsive interaction observed here between oxo species causes a destabilization of adsorbed *O species (ΔG*O-ΔG*OH becomes larger), thus reducing the free energy of the RDS (O-O bond formation from *O).
From Fig. 3D, it is also apparent that at the same *O coverage, the intrinsic activity of the RDS on rutile IrO2 is around 1 order of magnitude higher (as measured by TOF (O2 redox3− 1 s− 1)) and factor of 5 higher (as measured by TOF (1/4τ)) than in amorphous IrOx. We rationalize this difference in activity to the potential dependent binding energetics of *O. From the experimentally observed energetics of *O, the values of ΔG*O-ΔG*OH at *O (θ*O = 0) for rutile IrO2 is around 0.1 eV larger than that of amorphous IrOx (requiring an extra 0.1 V to form *O on rutile), and both increase with *O coverage at a slope of roughly 0.15 eV/coverage due to the similar Frumkin interaction strengths (Fig. 3E). As a result, at a constant coverage of *O, compared with amorphous IrOx, rutile IrO2 binds oxygen more weakly, which facilitates O-O bond formation from *O in the rate-determining step, leading to higher intrinsic activity per active state. On the other hand, at a constant potential, for example at 1.48 V, this binding energy difference decreases to around 0.05 eV because amorphous IrOx has a higher coverage of *O at this potential, thus weakening the *O binding more significantly. Therefore, the intrinsic activity at a constant potential is similar, and the order of magnitude greater geometric-area-normalized current density observed for the amorphous is mainly because it has a higher concentration of active states, as shown in Fig. 2D,E.
Design Principles For Highly Active Catalytic Materials For Water Oxidation
The potential (and coverage) dependence of the intrinsic water oxidation kinetics determined herein show the importance of the absorbate-absorbate interactions in controlling the kinetics of the OER reaction. The conventional approach for catalyst design has focussed on developing catalysts with optimal binding energetics (ΔG*O-ΔG*OH) corresponding to approximately 1.6 eV 24, 25, 44. This implicitly assumes that this value is coverage independent. Herein, we have clearly experimentally demonstrated the weakening of the *O binding energy with increasing coverage. This indicates that optimal binding energetics can also be obtained on surface sites that are strongly binding at low coverage ((ΔG*O-ΔG*OH <1.6 eV at θ*O ~ 0), but have strong repulsive interactions, which weaken *O binding energies at high coverage, thereby enabling these sites to achieve the optimal ΔG*O-ΔG*OH values under reaction conditions corresponding to higher coverage.
To include the influence of absorbate-absorbate interactions on ΔG*O-ΔG*OH values, and thus the intrinsic activity per state, we construct a new 3-dimensional volcano plot. This volcano plot can be seen as a deconvolution of the previous descriptor ∆GO - ∆GOH into two components (1) binding energetics on a surface with zero *O coverage, ∆GO - ∆GOH(θ*O = 0), and (2) interaction strength between *O species, following a Frumkin-isotherm-related equation [∆GO - ∆GOH = ∆GO - ∆GOH(θ*O = 0) + r*θ*O]. We first construct this plot assuming near saturated coverage of θ*O = 0.9. As shown in Fig. 4A, the optimal activity of the states can be achieved either (1) via an optimal *O binding at low coverage without *O interaction (i.e. ΔG*O-ΔG*OH ~1.6 eV at θ*O ~ 0, r = 0 eV), which is the top of conventional volcano plot; or (2) via a stronger *O binding than optimal and strong interaction (e.g. ΔG*O-ΔG*OH=1 eV, at θ*O ~ 0, r ~ 0.7 eV). In the latter case, although each state binds *O much stronger than optimal at zero coverage, the strong interaction strength results in weakening binding and optimum ΔG*O-ΔG*OH at high coverage.
We note that the applied potentials to achieve the same coverage are different for catalysts with different ∆GO - ∆GOH(θ*O = 0) and r values. To include both potential and coverage effects, we construct a volcano plot at a fixed overpotential of 250 mV (Fig. 4B, and supplementary Fig. 28). The steep part of the left side of the volcano (for example, when ∆GO - ∆GOH(θ*O = 0) < 1.18 eV in the curve of r = 0.15 eV) is where *O coverage is saturated, whereas the less steep part represents where ∆GO - ∆GOH(θ*O) is changing due to the Frumkin repulsion term. The Frumkin repulsion term has a higher impact on increasing the activity for a material with stronger *O binding energy, for a given interaction strength. This is evident from our data on iridium-based catalysts, where the absorbate-absorbate interaction increases the relative activity of the two materials by different degrees, although they have similar interaction parameter (~ 0.15 eV). This is because the stronger binding strength of *O on the IrOx surface results in a higher coverage for IrOx compared to IrO2 at this potential (Fig. 2D,E), and consequently higher repulsion between adsorbates. However, both show higher activity than the molecular Ir catalyst analysed previously, which have no adsorbate-adsorbate interaction.28 This observation is consistent with our previous suggestion that although the studied molecular Ir showed faster kinetics than amorphous IrOx at low overpotentials, it is less active than amorphous IrOx at higher overpotentials due to a lack of cooperative interaction between active states28. Similar analyses at overpotentials of 200 and 400 mV are also given in Supplementary Fig. 29. These results indicate that potential can alter the coverage of *O species, which in turn can weaken the binding energies for catalysts that bind oxygen too strongly in a conventional volcano plot, hence improving OER activity. This analysis elucidates the fundamental origin of the activity discrepancies between different Ir-based catalysts, including amorphous and crystalline rutile iridium oxides.