Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2019 Nov 4;116(47):23473–23479. doi: 10.1073/pnas.1901046116

Short O–O separation in layered oxide Na0.67CoO2 enables an ultrafast oxygen evolution reaction

Hao Wang a,b,c,d,1, Jinpeng Wu e,f,1, Andrei Dolocan a,b, Yutao Li a,b,2, Xujie Lü g, Nan Wu a,b, Kyusung Park a,b, Sen Xin a,b, Ming Lei c,d, Wanli Yang e, John B Goodenough a,b,2
PMCID: PMC6876228  PMID: 31685612

Significance

The development of a low-cost, stable, and more active electrocatalyst for the oxygen evolution reaction (OER) is critical for the practical storage of electric power in hydrogen gas produced by the electrolysis of water. We demonstrate a stable OER of higher rate at a lower voltage with a low-cost oxide that provides, as an alternative to the conventional reaction route, a faster route that is greatly enhanced by an unusually short O–O separation.

Keywords: air electrodes, water electrolysis, catalytic mechanisms, structure−property relationship

Abstract

The layered oxide Na0.67CoO2 with Na+ occupying trigonal prismatic sites between CoO2 layers exhibits a remarkably high room temperature oxygen evolution reaction (OER) activity in alkaline solution. The high activity is attributed to an unusually short O–O separation that favors formation of peroxide ions by O–O interactions followed by O2 evolution in preference to the conventional route through surface O–OH species. The dependence of the onset potential on the pH of the alkaline solution was found to be consistent with the loss of H+ ions from the surface oxygen to provide surface O that may either be attacked by solution OH or react with another O; a short O–O separation favors the latter route. The role of a strong hybridization of the O–2p and low-spin CoIII/CoIV π-bonding d states is also important; the OER on other CoIII/CoIV oxides is compared with that on Na0.67CoO2 as well as that on IrO2.


Highly active oxygen evolution reaction (OER) catalysts with a long-term stability are required to reduce the energy loss, increase the rate performance, and improve the cycling stability of different energy conversion and storage systems, particularly in electrochemical water electrolysis and rechargeable metal–air aqueous electrolyte batteries (14). The most active OER catalyst, IrO2, is expensive and shows a high overpotential of 0.3 V at 10 mA⋅cm−2; moreover, it is unstable at the applied potential in an alkaline electrolyte, which degrades its activity and limits its application (57).

Low-cost transition metal (TM) oxides are promising candidates for OER catalysts (817). AMO3 perovskites with controllable chemical compositions and electronic structures by substituting the cations on A and M sites have been extensively investigated as OER catalysts (1822). Some perovskite oxides (e.g., Ba0.5Sr0.5Co0.8Fe0.2O3-δ, Hg2Ru2O7, and CaCu3Fe4O12) exhibit comparable or even better OER catalytic activity than that of IrO2 (3, 5, 23). However, the obtained OER descriptor of eg ≈ 1 on the d-orbital manifold of Mn+ ions in AMO3 perovskites for their excellent OER activity is challenged by the good OER on 4d/5d TM oxide catalysts with zero antibonding eg electrons such as IrO2 and Hg2Ru2O7 (2427). In addition to the OER activity, the stability and the preparation condition of the catalysts are 2 other critical parameters for their large-scale application. For example, the crystalline Ba0.5Sr0.5Co0.8Fe0.2O3-δ phase becomes amorphous after OER testing in an alkaline electrolyte, and the stable Hg2Ru2O7 and CaCu3Fe4O12 catalysts with high valence Ru5+ and Fe4+ ions can only be prepared at an extremely high pressure of 6 and 15 GPa, respectively (23). The instability of TM oxides originates from cations that can dissolve into the alkaline solution and from lattice oxygen loss during the OER (28). Increasing the Mn+–O2– bond strength of oxides helps to improve their stability. Therefore, developing an easily prepared, efficient, and durable OER catalyst based on Earth-abundant elements is still a challenge.

We have recently studied the onset potentials and OER activities of 2 cubic perovskites, CaCoO3 and SrCoO3, prepared under high pressure (29); both have CoIV ions having similar intermediate spin states t4σ*1, but CaCoO3 had a significantly shorter CoIV–O bond (1.87 Å) and showed a higher OER activity than SrCoO3 as a result of its reduced lattice parameter. After surface deprotonation at a charging potential Vch = Von, where Von was the onset potential, the surface CoIV–O were attacked by solution OH,

CoIVOe-+OH-=CoIVOOH-, [1]

but the competitive reaction

2CoIVO+OH-=CoIV(O2)2+CoIVOH- [2]

followed by

2CoIV(O2)22e-=2CoIVO+O2 [3]

was much stronger on CaCoO3 because of its much smaller lattice parameter, which reduced the surface O–O separation. The O + O = (O2)2– reaction is faster the shorter the O–O separation of the catalyst. In the conventional reaction route, the O ion is attacked by a solution OH to form OOH, and the OOH + OH = (O2)2– + H2O; this mechanism is independent of the O–O separation.

These findings recommended to us a search for stable oxides with a shorter surface oxygen separation that can be prepared at ambient pressure. Therefore, we prepared a metallic layered oxide Na0.67CoO2 with low-spin CoIII/IV ions (CoIII: π*6σ*0; CoIV: π*5σ*0) and a much shorter O–O separation than on CaCoO3 to study further the influence of zeta potential, surface oxygen separation, and the pH of the aqueous medium on the onset potential and OER activity. Metallic, low-spin Na0.67CoO2 has a strong CoIII/IV–O interaction as a result of empty σ-antibond eg orbitals and itinerant π-bonding electrons, which reduces the charge transfer resistance during the OER; Na0.67CoO2 has a zero zeta potential at pH = 4 and shows a pH-dependent onset potential for the OER consistent with the potential for the CoOH – e + OH = CoO + H2O reaction.

Results

The Crystal and Electronic Structure of Na0.67CoO2.

Synchrotron powder X-ray diffraction (XRD) and powder X-ray diffraction patterns of as-prepared Na0.67CoO2 (Fig. 1 and SI Appendix, Fig. S1) confirm the layered structure of Na0.67CoO2. Two-dimensional CoO2 layers with edge-sharing CoO6 octahedra in trigonal Na0.67CoO2 (space group: R-3c) are separated by sodium layers. The Na+ occupy 3 different trigonal prismatic sites, Na1 and Na2 (Na2a and Na2b) in SI Appendix, Table S1; each shares 6 edges and 2 triangular faces along the c axis with the CoO6 octahedra vertically above and below. The Na1 ions are energetically less favorable because of the coulombic repulsions from 2 Co ions in face-sharing octahedra. The Co ions of Na0.67CoO2 have 4 different positions; the distorted Co–O octahedra have a similar average Co–O bond distance, but one shortest O–O separation (2.30 Å; Fig. 1F) than that of cubic CaCoO3 (O–O: 2.64 Å) in which the CoIV have an intermediate spin state (t4σ*1). About 0.3 wt% Co3O4 impurity was found to exist in the sample by refining the synchrotron data. The Na0.67CoO2 sample had an average particle size of 15 μm (SI Appendix, Fig. S3), and the energy dispersive spectroscopy (EDS) mapping revealed a uniform distribution of Na, Co, and O elements.

Fig. 1.

Fig. 1.

Crystal structure and magnetic and transport properties of Na0.67CoO2. (A) Observed, calculated patterns, and their difference for the Rietveld refinement of the synchrotron XRD of Na0.67CoO2; (Inset) its crystal structure (pink, Na; blue, Co; red, O). (B and C) TEM images of Na0.67CoO2 before and after OER measurements. (D) Temperature dependence of resistivity and magnetization. (E) Electronic spin states of CoIV ions and schematic band diagrams of Na0.67CoO2. (FI) The O–O bonds shorter than (2.4 to 2.64) Å are highlighted with thick red lines.

The electronic conductivity and magnetic properties of Na0.67CoO2 were investigated to determine the d-electron configuration of the CoIII/IV ions (Fig. 1D). The temperature dependence of resistivity of Na0.67CoO2 shows a metallic behavior down to 2 K, with a large residual resistance ratio; a nonlinear magnetic susceptibility curve from 2 to 300 K is also consistent with itinerant d electrons. A spin-polarized first-principle calculation was performed, and the density of states of Co ions from Na0.67CoO2 is shown in Fig. 1E. The conduction band of Na0.67CoO2, which contains one spin component of t2g orbitals, is not full, indicating a metallic conductivity of Na0.67CoO2. The higher electronic conductivity of Na0.67CoO2 reduces the charge transfer resistance (Rct) and the ohmic potential drop of the OER; the Rct of Na0.67CoO2 is one order of magnitude smaller than that of the electronic insulator Co3O4, which is above 1.6 V (SI Appendix, Fig. S4).

Soft X-ray absorption spectroscopy (sXAS), which is quite sensitive to the TM 3d states because of the strong dipole-allowed 2p–3d (L edge) excitations (3033), was performed to explore the valence and spin states of the Co ions of Na0.67CoO2 (Fig. 2). With a probe depth of 10 nm, the total electron yield (TEY) mode of sXAS is quite an effective surface-sensitive probe, as shown in Fig. 2 A and B. Due to the core-hole spin–orbit coupling, the Co L edge in Fig. 2A is well separated into 2 branches, that is, L3 (between 774 and 787 eV) and L2 (between 788 and 799 eV). The branching ratio of the L3 and L2 edges, which is greatly affected by the spin–orbit coupling and especially by the 2p–3d multiplet effects, can be utilized to deduce the spin state of Co (34, 35). Quantitatively, for the statistical ratio of the integrated intensity of 2 edges, namely IL3/(IL3 + IL2), the high-spin state has a higher one than the low-spin state. Respectively for the samples before and after catalysis reaction, the statistical branching ratio of the Co L inverse partial fluorescence yield (iPFY) is 0.689 ± 0.005 and 0.686 ± 0.005. Based on the atomic multiplet calculations (36, 37), the standard spectra of LS Co within the octahedral (CoO6) structure is achieved as shown by dashed lines in Fig. 2 A and B, and that of HS is done as shown in SI Appendix, Fig. S6. While the calculated spectrum of LS Co (either Co3+ or Co4+) has a branching ratio of 0.698 ± 0.005, very close to the experimental Co L iPFY, that of HS has a much higher one of 0.780 ± 0.005, suggesting that Na0.67CoO2 has a low-spin state (CoIII: t6e0, CoIV: t5e0). The density of states of Co ions from Na0.67CoO2, which is shown in Fig. 1E, also indicates low-spin CoIII/IV ions. The valence state of the Co ions can be inferred by the Co L3 peak position. As shown in Fig. 2B, the peak A located at 777.6 eV, which is 0.2 eV from the standard Co3+ (777.4 eV) and 0.4 eV from the standard Co4+ (778.0 eV). This position indicates that the valence of the Co ions is about 3.3+, consistent with the expected stoichiometry as prepared.

Fig. 2.

Fig. 2.

The absorption profiles of CoIII and CoIV in Na0.67CoO2 probed by sXAS and mRIXS. (A and B) The sXAS TEY of (A) Co L edge and (B) magnified L3 edge obtained with a surface probe having a probe depth of 10 nm. (C and D) The mRIXS iPFY of (C) Co L edge and (D) magnified L3 edge by a bulk probe with a probe depth of 100 nm.

In addition, we demonstrate the bulk information as a supplement. Resulting from the self-absorption and saturation effects (38, 39), the bulk probe of sXAS, that is, total fluorescence yield mode with a probe depth of 100 nm (40), provided a noisy lineshape and an unreliable L2/L3 intensity ratio as shown in SI Appendix, Fig. S5. The iPFY is theoretically an undistorted absorption profile that can be extracted from the map of resonant inelastic X-ray scattering (mRIXS). In this work, we performed the Co L mRIXS measurement on Na0.67CoO2, and achieved the iPFY for characterizing the bulk Co status. (For the convenience of the reader, we introduce how to extract the Co L iPFY from the mRIXS in SI Appendix, Fig. S5.) As shown in Fig. 2 C and D, the Co L iPFY spectra demonstrate a consistent lineshape with the TEY spectra, indicating that Co ions in the bulk of Na0.67CoO2 present the same low-spin electronic and 3.3+ valence state as those on the surface.

The OER Activity of Na0.67CoO2.

The OER performance of Na0.67CoO2 was compared with that of rutile IrO2, spinel Co3O4, and layered Co(OH)2 (Fig. 3). The current densities of all of the samples were normalized to the electrochemically active surface area to exclude geometric effects (SI Appendix, Fig. S7 and Table S2). The catalysts with different particle size have similar OER activities when the current density is normalized to the electrochemical surface area or the surface area confirmed by Brunauer–Emmett–Teller measurement (41). Na0.67CoO2 with an onset potential of 1.5 V vs. a reversible hydrogen electrode (RHE) had the smallest overpotential (0.29 V) at 10 mA cm−2 and the highest current density at voltages above 1.6 V; the layered Co(OH)2 and spinel Co3O4 exhibited a negligible catalytic current density compared with Na0.67CoO2. The smallest Tafel slope of 39 mV⋅dec−1 (Fig. 3C) for Na0.67CoO2 also indicates its excellent OER kinetics. The layered Li1-xCoO2 has a structure and CoIII/IV–O bond similar to that in Na0.67CoO2, but the Li+ are in octahedral sites and no O–O separation is reduced; it has a much higher onset potential and a smaller OER current density than Na0.67CoO2 (6).

Fig. 3.

Fig. 3.

OER performance of Na0.67CoO2, IrO2, Co(OH)2, and Co3O4. (A) Linear sweep voltammograms at 1,600 revolutions per minute. in 0.1 M KOH. (B) The overpotential at 10 mA⋅cm−2 (dashed line). (C) Tafel plots. (D) The chronoamperometric curves in an O2-saturated 0.1 M KOH electrolyte at 1.6 V vs. RHE and the CV curves of first, 500th, and 1,000th cycles (Inset).

The Surface Charge Density of Na0.67CoO2.

The mean surface charge density of Na0.67CoO2 (Fig. 4A) was evaluated by its zeta potential in water with different pH. Na0.67CoO2 has a zero zeta potential at pH = 4; the strong covalence of the Co III/IV–O bond of Na0.67CoO2 makes it more acidic than the spinel Co3O4, which has a weaker CoII/III–O bond and a high zero-zeta potential at pH = 7.5. IrO2 with a strong IrIV–O bond has a similar zeta potential curve to that of Na0.67CoO2, indicating analogous acidity of the surface states and pH influence on their OER.

Fig. 4.

Fig. 4.

The pH-dependent OER behavior of Na0.67CoO2. (A) Zeta potential of the catalysts. (B and C) CV measurements of Na0.67CoO2 in O2-saturated KOH with pH 12.5 to 14. Inset shows the enlarged CV part from 1.3 to 1.6 V. (D) The slope change of the linear CV curves at voltages above 1.7 V in B and C and SI Appendix, Fig. S8 with different pH.

The onset potential and activity of Na0.67CoO2 at different pH are compared with those of Co3O4, Co(OH)2 and IrO2 in Fig. 4 and SI Appendix, Fig. S8. The oxidation voltage of surface CoIII to CoIV and the onset potential of Na0.67CoO2 were reduced with increasing pH (Fig. 4B) because of the easier deprotonation process at higher pH; the activities of Na0.67CoO2 and IrO2 show a similar pH-dependent behavior on the RHE scale, and their OER currents increase significantly at high pH; however, both Co3O4 and Co(OH)2 have no significant OER current increase. The slope of all CV curves of these catalysts at voltages above 1.6 V increases exponentially with pH (Fig. 4D), and the most active Na0.67CoO2 has the biggest slope of all of the catalysts at different pH. Both Na0.67CoO2 and IrO2 have a much larger slope change than Co3O4 and Co(OH)2, indicating different OER mechanisms in these catalysts.

The OER Stability of Na0.67CoO2.

Na0.67CoO2 shows an excellent OER stability in an O2-saturated electrolyte with pH = 13 (Fig. 3D). More than 90% of its initial current density after 20,000 s was retained; and, after 1,000 cyclic voltammetry (CV) cycles, Na0.67CoO2 showed almost the same OER activity. The transmission electron microscopy (TEM) image (Fig. 1C) and XRD (SI Appendix, Fig. S1) results confirmed the same crystalline structure of bulk and surface Na0.67CoO2 before and after OER testing. The X-ray photoelectron spectroscopy (XPS) spectra of fresh Na0.67CoO2, Na0.67CoO2 soaked in KOH for a week, and Na0.67CoO2 after 1,000 CV cycles are shown in SI Appendix, Fig. S9; the Na, Co, and O peaks of Na0.67CoO2 after cycling are much weaker because of the covering Nafion binder on the particle surface; all of these XPS peaks retain the same positions, verifying the good stability of Na0.67CoO2 during OER. Na0.67CoO2 shows a stable peak position in sXAS, and the valence state of Co sXAS remains unchanged in the Na0.67CoO2 powder before and after OER (Fig. 2). The strong Co–O bonds of Na0.67CoO2 increase its stability in alkaline solution.

The Surface of Na0.67CoO2 after OER.

Because surface oxygens of Na0.67CoO2 participate in the OER, bulk lattice oxygen can diffuse to the surface oxygen vacancies after O2 gas release and capture a proton from solution before the solution OH enters, and then the vacancies will be generated on these oxygens. The Co and O ions of a minimum 14-nm-thick surface of perovskite SrCoO3-x have been reported to participate in the OER (42).

Time-of-flight secondary ion mass spectrometry (TOF-SIMS), which is an ultrahigh elemental and surface sensitive technique, was employed to study whether any chemical composition change occurs on the surface of Na0.67CoO2 before and after OER testing. Given the destructive nature of TOF-SIMS, all ionized fragments detected imply chemical bonding between the fragment elements prior to sputtering (43). TOF-SIMS depth profiling and high-resolution mapping were used to show the presence of CoOH and CoO2H on the surface of the Na0.67CoO2 particles following OER (Fig. 5 and SI Appendix, Fig. S10). Due to the naturally high surface corrugation of Na0.67CoO2, both CoOH and CoO2H secondary ion depth profiles were normalized by the corresponding Co profile in each sample to account for the topography changes between the Na0.67CoO2 surfaces before and after OER. As a proxy for bulk Na0.67CoO2, the Co signal was selected for normalization. Finally, we used the ratio between the Co–normalized CoOH and CoO2H profiles before and after OER to demonstrate their surface localization after the OER in a ∼70-nm layer; the peak position of this ratio in Fig. 5 A and B provided the localization. In comparison, the CoO/Co profile appears less enhanced at the surface, which suggests the OER produces only a limited amount of CoO at the Na0.67CoO2 surface (Fig. 5C). However, given the natural fragmentation of Na0.67CoO2 upon sputtering, the CoO signal could also be used as a marker for the bulk Na0.67CoO2. As such, TOF-SIMS high-resolution mapping of CoO and CoOH secondary ion fragments was performed to show directly the formation of CoOH at the surface of Na0.67CoO2 (Fig. 5D). Indeed, deep sputtering of the Na0.67CoO2 particles confirms the continuous presence of CoOH at their surface, as presented in Fig. 5D in the dual color overlay as a function of depth, that is, of Cs+ sputtering time. The Na/Co ratio of Na0.67CoO2 was confirmed, by inductively coupled plasma mass spectrometry, not to change from before to after OER testing (SI Appendix, Table S3).

Fig. 5.

Fig. 5.

TOF-SIMS depth profiling and high-resolution mapping of Na0.67CoO2 particles before and after OER. (AC) TOF-SIMS depth profiling of Na0.67CoO2 before and after cycling. The intermediates CoOH, CoO2H, and CoO increase after cycling. (D) Dual color overlays of high-resolution maps of the CoO and CoOH secondary ion signals demonstrating that CoOH mainly exists on the surface of the catalyst particles.

Discussion

In air, an exposed surface cation M of an oxide, especially one with a strong octahedral site preference energy, attracts water to complete its oxygen coordination; the hydrogen atoms of the adsorbed water H2O are dispersed over the surface oxygen to create surface M–OH. In alkaline solution, an exposed surface cation attracts solution OH when it is oxidized during charge. Continuing removal of electrons from the oxide during charge in KOH solution induces the reaction (Fig. 6A)

M-OH--e-+OH-=M-O-+H2O. [4]

This reaction occurs at a critical potential VcVon, where Von is the onset potential for the OER to occur with continuing charge; the resulting MO may be attacked by the solution OH,

M-O--e-+OH-=M-OOH-, [5]

and the subsequent OER activity then depends on the relative rates of removal of the H+,

M-OOH--2e-+2OH-=MOH-+H2O+O2, [6]

and displacement of the O2H by a solution OH. An O2H can be an unwanted by-product of reaction 5.

Fig. 6.

Fig. 6.

Competing OER mechanisms on Na0.67CoO2. OER mechanism with surface lattice oxygen activated for OER to form peroxide either (A) in a CoO2 layer or (B) between 2 neighboring CoO2 layers.

Alternatively, 2 neighboring surface O may react with one another (Fig. 6B),

2MO--e-+OH-=M(O2)-+MOH- [7]

followed by

M(O2)--e-+OH-=MOH-+O2. [8]

The rate of reaction 7 competes with that of reaction 5; it increases strongly with decreasing surface O–O separation. Therefore, the activity of the OER above the onset potential depends strongly on the surface O–O separation. The observation of an ultrafast OER in Na0.67CoO2 with an unusually short O–O separation demonstrates this dependence. A layer oxide NaxCoO2 with x = 0.52, 0.65, and 0.75 has a different crystal structure (space group: P63/mmc) with our Na0.67CoO2 (R-3c); all of these 3 materials show almost the same OER activity after cycling. Our Na0.67CoO2 shows a much higher OER activity than the reported NaxCoO2; the overpotential at 10 mA⋅cm−2 of our Na0.67CoO2 is 0.29 V, which is much smaller than their NaxCoO2 (0.45 to 0.47 V) (44). In NaxCoO2 with space group P63/mmc, for example, Na0.65CoO2, Na and Co ions occupy 2 and 1 different positions, respectively; however, there are 3 Na and 4 Co positions in Na0.67CoO2 with space group R-3c. The O–O separation in Na0.65CoO2 is 2.57 Å, while the O–O separation of our Na0.67CoO2 is much smaller (the shortest O–O is 2.30 Å); the large O–O separation difference is caused by the different crystal structure and Na ordering. Their results also well support our conclusion that short O–O separation determines the OER activity. The relationship between the catalytic performance and the O–O bond length in NaxCoO2 and LixCoO2 with different Na+ and Li+ proportion according to the previous report is shown in SI Appendix, Fig. S2. The shorter the O–O bond length in these oxides, the smaller the overpotentials at 5 mA⋅cm−2, which confirms the key role of the short O–O separation in the OER performance of the Na0.67CoO2.

Reaction 4, which sets the onset potential, depends on the zeta potential of the oxide and the pH of the solution. Reaction 4 is favored the higher the pH of the solution and the greater the acidity of the oxide. The stronger the M–O bond, the more acidic is the oxide. The existence of itinerant electrons in π-bonding orbitals of d-wave symmetry not only lowers the resistance to the OER, but also testifies to a strong O-2p hybridization in the π-bonding as well as σ-bonding orbitals of d-wave symmetry.

Summary

The excellent OER activity of Na0.67CoO2 is the result of a short O–O separation that increases the rate of reaction 7 relative to reaction 5 and demonstrates the superior rate capability of reaction 7. The stability of the catalyst is the result of strong Co–O σ-bonding with the CoIV configuration on π*5σ*0.

Materials and Methods

Na0.67CoO2 was prepared by a typical solid-state reaction with analytical grade Na2CO3 and Co3O4 as raw materials. First of all, the mixture of Na2CO3 and Co3O4 in a stoichiometric ratio with an excess of 5 mol% Na2CO3 was thoroughly ground in agate bowls, pressed into pellets, and sintered at 700 °C for 2 d, 900 °C for 15 h, and 950 °C for 10 h at a heating rate of 3 °C⋅min−1 with intermediate grinding. These pellets, after sintering, were crushed and powdered to obtain a fine particle size. For comparison, commercial IrO2, Co(OH)2, and Co3O4 were purchased from Alfa Aesar and used without further purification.

Supplementary Material

Supplementary File

Acknowledgments

H.W. thanks the China Scholarship Council for the opportunity to work in Texas. This work was supported by the Department of Energy (DOE), Office of Energy Efficiency and Renewable Energy, under Award DE–EE0007762, and the Robert A. Welch Foundation of Houston, TX. This research used resources of the Advanced Light Source, which is a DOE Office of Science User Facility under Contract DE-AC02-05CH11231 used synchrotron resources of the Advanced Photon Source (Sector 11 ID-C), a US Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02-06CH11357. Works at Stanford are supported by the Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division, under Contract DE-AC02-76SF00515.

Footnotes

The authors declare no competing interest.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1901046116/-/DCSupplemental.

References

  • 1.Chen D., Chen C., Baiyee Z. M., Shao Z., Ciucci F., Nonstoichiometric oxides as low-cost and highly-efficient oxygen reduction/evolution catalysts for low-temperature electrochemical devices. Chem. Rev. 115, 9869–9921 (2015). [DOI] [PubMed] [Google Scholar]
  • 2.Jiao Y., Zheng Y., Jaroniec M., Qiao S. Z., Design of electrocatalysts for oxygen- and hydrogen-involving energy conversion reactions. Chem. Soc. Rev. 44, 2060–2086 (2015). [DOI] [PubMed] [Google Scholar]
  • 3.Suntivich J., May K. J., Gasteiger H. A., Goodenough J. B., Shao-Horn Y., A perovskite oxide optimized for oxygen evolution catalysis from molecular orbital principles. Science 334, 1383–1385 (2011). [DOI] [PubMed] [Google Scholar]
  • 4.Goodenough J. B., Manoharan R., Paranthaman M., Surface protonation and electrochemical activity of oxides in aqueous solution. J. Am. Chem. Soc. 112, 2076–2082 (1990). [Google Scholar]
  • 5.Mefford J. T., et al. , Water electrolysis on La 1-xSr xCoO 3-δ perovskite electrocatalysts. Nat. Commun. 7, 11053 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Maiyalagan T., Jarvis K. A., Therese S., Ferreira P. J., Manthiram A., Spinel-type lithium cobalt oxide as a bifunctional electrocatalyst for the oxygen evolution and oxygen reduction reactions. Nat. Commun. 5, 3949 (2014). [DOI] [PubMed] [Google Scholar]
  • 7.Rasiyah P., Tseung A. C. C., The role of the lower metal oxide/higher metal oxide couple in oxygen evolution reactions. J. Electrochem. Soc. 131, 803–808 (1984). [Google Scholar]
  • 8.Subbaraman R., et al. , Trends in activity for the water electrolyser reactions on 3d M(Ni,Co,Fe,Mn) hydr(oxy)oxide catalysts. Nat. Mater. 11, 550–557 (2012). [DOI] [PubMed] [Google Scholar]
  • 9.Han L., Dong S., Wang E., Transition-metal (Co, Ni, and Fe)-based electrocatalysts for the water oxidation reaction. Adv. Mater. 28, 9266–9291 (2016). [DOI] [PubMed] [Google Scholar]
  • 10.Zheng X., et al. , Theory-driven design of high-valence metal sites for water oxidation confirmed using in situ soft X-ray absorption. Nat. Chem. 10, 149–154 (2018). [DOI] [PubMed] [Google Scholar]
  • 11.Lu Z., et al. , Identifying the active surfaces of electrochemically tuned LiCoO2 for oxygen evolution reaction. J. Am. Chem. Soc. 139, 6270–6276 (2017). [DOI] [PubMed] [Google Scholar]
  • 12.Oh H. S., et al. , Electrochemical catalyst-support effects and their stabilizing role for IrOx nanoparticle catalysts during the oxygen evolution reaction. J. Am. Chem. Soc. 138, 12552–12563 (2016). [DOI] [PubMed] [Google Scholar]
  • 13.Tao H. B., et al. , Identification of surface reactivity descriptor for transition metal oxides in oxygen evolution reaction. J. Am. Chem. Soc. 138, 9978–9985 (2016). [DOI] [PubMed] [Google Scholar]
  • 14.Yan D., et al. , Defect chemistry of nonprecious-metal electrocatalysts for oxygen reactions. Adv. Mater. 29, 1606459 (2017). [DOI] [PubMed] [Google Scholar]
  • 15.Grimaud A., et al. , Activation of surface oxygen sites on an iridium-based model catalyst for the oxygen evolution reaction. Nat. Energy 2, 16189 (2016). [Google Scholar]
  • 16.Song F., et al. , Transition metal oxides as electrocatalysts for the oxygen evolution reaction in alkaline solutions: An application-inspired renaissance. J. Am. Chem. Soc. 140, 7748–7759 (2018). [DOI] [PubMed] [Google Scholar]
  • 17.Lu Z., et al. , Electrochemical tuning of layered lithium transition metal oxides for improvement of oxygen evolution reaction. Nat. Commun. 5, 4345 (2014). [DOI] [PubMed] [Google Scholar]
  • 18.Hong W. T., et al. , Toward the rational design of non-precious transition metal oxides for oxygen electrocatalysis. Energy Environ. Sci. 8, 1404–1427 (2015). [Google Scholar]
  • 19.Lee J. G., et al. , A new family of perovskite catalysts for oxygen-evolution reaction in alkaline media: BaNiO3 and BaNi0.83O2.5. J. Am. Chem. Soc. 138, 3541–3547 (2016). [DOI] [PubMed] [Google Scholar]
  • 20.Petrie J. R., et al. , Enhanced bifunctional oxygen catalysis in strained LaNiO3 perovskites. J. Am. Chem. Soc. 138, 2488–2491 (2016). [DOI] [PubMed] [Google Scholar]
  • 21.Fabbri E., et al. , Dynamic surface self-reconstruction is the key of highly active perovskite nano-electrocatalysts for water splitting. Nat. Mater. 16, 925–931 (2017). [DOI] [PubMed] [Google Scholar]
  • 22.Yamada I., et al. , Bifunctional oxygen reaction catalysis of quadruple manganese perovskites. Adv. Mater. 29, 1603004 (2017). [DOI] [PubMed] [Google Scholar]
  • 23.Yagi S., et al. , Covalency-reinforced oxygen evolution reaction catalyst. Nat. Commun. 6, 8249 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Hirai S., et al. , Non-Fermi liquids as highly active oxygen evolution reaction catalysts. Adv. Sci. (Weinh.) 4, 1700176 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Rao R. R., et al. , Towards identifying the active sites on RuO2 (110) in catalyzing oxygen evolution. Energy Environ. Sci. 10, 2626–2637 (2017). [Google Scholar]
  • 26.Kuo D. Y., et al. , Influence of surface adsorption on the oxygen evolution reaction on IrO2(110). J. Am. Chem. Soc. 139, 3473–3479 (2017). [DOI] [PubMed] [Google Scholar]
  • 27.Kim Y. T., et al. , Balancing activity, stability and conductivity of nanoporous core-shell iridium/iridium oxide oxygen evolution catalysts. Nat. Commun. 8, 1449 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Binninger T., et al. , Thermodynamic explanation of the universal correlation between oxygen evolution activity and corrosion of oxide catalysts. Sci. Rep. 5, 12167 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Li X., et al. , Exceptional oxygen evolution reactivities on CaCoO3 and SrCoO3. Sci. Adv. 5, eaav6262 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Lin F., et al. , Synchrotron X-ray analytical techniques for studying materials electrochemistry in rechargeable batteries. Chem. Rev. 117, 13123–13186 (2017). [DOI] [PubMed] [Google Scholar]
  • 31.Yang W., et al. , Key electronic states in lithium battery materials probed by soft X-ray spectroscopy. J. Electron Spectrosc. Relat. Phenom. 190, 64–74 (2013). [Google Scholar]
  • 32.Zhang H., et al. , Developing soft X-ray spectroscopy for in situ characterization of nanocatalysts in catalytic reactions. J. Electron Spectrosc. Relat. Phenom. 197, 118–123 (2014). [Google Scholar]
  • 33.Wu J., et al. , Modification of transition-metal redox by interstitial water in hexacyanometalate electrodes for sodium-ion batteries. J. Am. Chem. Soc. 139, 18358–18364 (2017). [DOI] [PubMed] [Google Scholar]
  • 34.De Groot F., Multiplet effects in X-ray spectroscopy. Coord. Chem. Rev. 249, 31–63 (2005). [Google Scholar]
  • 35.Thole B. T., van der Laan G., Branching ratio in x-ray absorption spectroscopy. Phys. Rev. B Condens. Matter 38, 3158–3171 (1988). [DOI] [PubMed] [Google Scholar]
  • 36.Chen J. M., et al. , A complete high-to-low spin state transition of trivalent cobalt ion in octahedral symmetry in SrCo0.5Ru0.5O3-δ. J. Am. Chem. Soc. 136, 1514–1519 (2014). [DOI] [PubMed] [Google Scholar]
  • 37.Li Q., et al. , Quantitative probe of the transition metal redox in battery electrodes through soft X-ray absorption spectroscopy. J. Phys. D Appl. Phys. 49, 413003 (2016). [Google Scholar]
  • 38.Achkar A. J., et al. , Bulk sensitive X-ray absorption spectroscopy free of self-absorption effects. Phys. Rev. B Condens. Matter Mater. Phys. 83, 081106 (2011). [Google Scholar]
  • 39.Achkar A. J., Regier T. Z., Monkman E. J., Shen K. M., Hawthorn D. G., Determination of total x-ray absorption coefficient using non-resonant x-ray emission. Sci. Rep. 1, 182 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Qiao R., et al. , Distinct solid-electrolyte-interphases on Sn (100) and (001) electrodes studied by soft X-ray spectroscopy. Adv. Mater. Interfaces 1, 1300115 (2014). [Google Scholar]
  • 41.Sun S., Li H., Xu Z. J., Impact of surface area in evaluation of catalyst activity. Joule 2, 1024–1027 (2018). [Google Scholar]
  • 42.Grimaud A., et al. , Activating lattice oxygen redox reactions in metal oxides to catalyse oxygen evolution. Nat. Chem. 9, 457–465 (2017). [DOI] [PubMed] [Google Scholar]
  • 43.Chou H., Ismach A., Ghosh R., Ruoff R. S., Dolocan A., Revealing the planar chemistry of two-dimensional heterostructures at the atomic level. Nat. Commun. 6, 7482 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Ji H., et al. , Investigating the origin of Co(IV)’s high electrocatalytic activity in the oxygen evolution reaction at a NaxCoO2 interface. Mater. Res. Bull. 95, 285–291 (2017). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES