Fully Oxidized Ni–Fe Layered Double Hydroxide with 100% Exposed

Jun 3, 2019 - The first one takes all the metal ions as catalytically active sites (TOFTM). ... indicates that Fe–O bonds contract from 1.99 to 1.88...
0 downloads 0 Views 2MB Size
Letter Cite This: ACS Catal. 2019, 9, 6027−6032

pubs.acs.org/acscatalysis

Fully Oxidized Ni−Fe Layered Double Hydroxide with 100% Exposed Active Sites for Catalyzing Oxygen Evolution Reaction Chunguang Kuai,†,‡,¶ Yan Zhang,†,§,¶ Deyao Wu,† Dimosthenis Sokaras,*,§ Linqin Mu,‡ Stephanie Spence,‡ Dennis Nordlund,§ Feng Lin,*,‡ and Xi-Wen Du*,† †

Institute of New-Energy Materials, School of Materials Science and Engineering, Tianjin University, Tianjin 300072, China Department of Chemistry, Virginia Tech, Blacksburg, Virginia 24061, United States § Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, Menlo Park, California 94025, United States Downloaded via CALIFORNIA INST OF TECHNOLOGY on July 21, 2019 at 02:15:05 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: Ni−Fe layered double hydroxides (LDHs) are promising for catalyzing the oxygen evolution reaction (OER) in alkaline media. However, the OER mechanism is highly debated, partially because of the lack of an ideal catalyst with 100% exposed active sites for unambiguous characterization. Herein, we develop an alcohol intercalation method to prepare ultrathin Ni−Fe LDH with a 1/3 unit-cell thickness and 100% exposed active sites. The ultrathin LDH catalyst exhibits an intrinsic activity similar to the bulk LDH and allows a direct and reliable characterization of the catalyst without any interference from “bulk” inactive species. Operando synchrotron X-ray analysis indicates that the metallic ions in ultrathin Ni−Fe LDH are fully oxidized into tetravalence states at low applied potentials and that the OER occurs on the tetravalent Ni and Fe ions following a decoupled proton/electron mechanism. Our findings demonstrate that a full oxidization of metal ions is crucial for highly active NiFe LDHs and that it can be accomplished by engineering ultrathin nanostructures. KEYWORDS: oxygen evolution reaction, catalysts, operando analysis, synchrotron X-ray absorption spectroscopy, catalytic mechanism, layer double hydroxide, active site

S

specific, their probing depth provides cumulative information from the entire studied material; namely, the detected signals may correspond to both the active ions at the surface and the nonactive ions at the catalyst subsurface and bulk. This may lead to an ambiguous determination of both the metal ions’ oxidation state in the OER active range and the catalytic mechanism. One possible solution is to engineer ultrathin catalysts (e.g., NiFe LDHs with single or double layer), which can expose all the metallic ions on the surface for operando XAS analysis. Previous works demonstrated that the thickness of NiFe LDH could be reduced via anion-exchange by weakening the interlayer van der Waals force.23,24 However, such a method, besides being time-consuming, also reduces the number of active sites because of the adsorption of residual exchanged anions on the catalyst surface.23 Therefore, it remains a challenge to synthesize few-layer NiFe LDHs for mechanistic study.

olar, wind, and hydro are among the most promising sustainable energy resources for substituting traditional fossil-fuel energy.1 To overcome their intermittency, energy storage in chemical molecules, such as hydrogen,2 hydrocarbons, alcohols,3,4 and ammonium,5−8 is imperative. Most of these chemical processes involve electrochemical reduction of small molecules such as H2O, CO2, and N2, with the water oxidation being a counter reaction through the oxygen evolution reaction (OER). Since the OER requires a fourelectron transfer, it is known to be very sluggish,9 and therefore, it is meaningful to develop scalable active water oxidation catalysts with optimized properties. Ni−Fe oxides and hydroxides are among the most efficient and earth-abundant OER catalysts in alkaline solution, with overpotentials of only 250−300 mV at a 10 mA/cm2 current density.10−14 Despite their high performance, the catalytic mechanism of Ni−Fe-based catalysts remains highly controversial because of the lack of an ideal catalyst with 100% exposed active sites for unambiguous investigation. Operando synchrotron X-ray absorption spectroscopy (XAS) has been widely used to study the local chemical environment, the oxidation states, and the catalytic behavior of OER catalysts.15−22 Although these techniques are element© 2019 American Chemical Society

Received: May 10, 2019 Revised: June 1, 2019 Published: June 3, 2019 6027

DOI: 10.1021/acscatal.9b01935 ACS Catal. 2019, 9, 6027−6032

Letter

ACS Catalysis

photoelectron spectroscopy (XPS) results illuminate that Ni/ Fe ratios in both samples are almost the same (Figures S1a,b) and that the iron valence state in ultrathin LDH is higher than that of the bulk counterpart (Figures S1c,d). To understand the electronic structure of the ultrathin LDH, we first investigate the starting materials by ex situ XAS. For the Ni cations, no obvious difference is observed between ultrathin and bulk LDHs in both K-edge and L-edge spectra (Figure 2a−c). In contrast, the Fe K-edge position of ultrathin

Herein, we develop an alcohol intercalation method to prepare ultrathin Ni−Fe LDH with a 1/3 unit-cell thickness (Scheme S1). The resulting catalyst achieves overpotentials of 210 and 320 mV at 10 and 500 mA/cm2, respectively, outperforming all NiFe LDHs reported in the literature.10,23,25,26 The turnover frequency (TOF) of ultrathin Ni−Fe LDH is similar to that of the bulk counterpart, indicating the same OER mechanism for both cases. The ultrathin nature of the catalyst (1/3 unit-cell) ensures that all its metal sites are at the surface, and hence, it has 100% exposed active sites. Therefore, this structure enables unambiguous XAS studies for the catalytic mechanism of mixed NiFe catalysts without the “bulk” interference. Our findings here show that ultrathin LDH can be fully oxidized under low OER potentials and that fully oxidized metal ions can boost the OER, which proceeds on NiFe LDH in a decoupled proton/electron way. The as-prepared ultrathin NiFe LDH exhibits a hexagonal morphology (Figure 1a) and a thickness of ∼1.2 nm (Figure

Figure 2. XAS results for the as-prepared samples. (a) Main edge of Ni K-edge, (b) zoomed in pre-edge region of Ni K-edge, (c) soft XAS of Ni L-edge in total electron yield(TEY) mode, (d) main edge of Fe K-edge, (e) zoomed in pre-edge region of Fe K-edge, and (f) soft XAS of Fe L-edge TEY mode. Ni(OH)2, FeCl2, and Fe2O3 were used as the reference samples for Ni2+, Fe2+, and Fe3+ ions, respectively. Figure 1. Transmission electron microscope images of (a) ultrathin and (b) bulk LDHs. Atomic force microscope images and thickness profiles of (c) ultrathin and (d) bulk LDHs, (e) XRD patterns for ultrathin and bulk LDHs. The standard α-Ni(OH)2 was extracted from JCPDS no. 38-0715. (f) Raman spectra of ultrathin and bulk LDHs.

LDH is 1.38 eV higher than that of bulk LDH (Figure 2d), which may arise from different Fe oxidation state or Fe−O covalency.37 The Fe oxidization state could be determined by analyzing the pre-edge region of K-edge (Figure 2e) and Ledge (Figure 2f) spectra. As shown in Figure 2e, two wellseparated peaks are observed for ultrathin LDH, which is typical for Fe3+ in the octahedral coordination (Oh).38 Comparatively, bulk LDH displays one wide peak, which may originate from the superposition of both Fe3+ and Fe2+ peaks. This hypothesis is verified by L-edge spectra (Figure 2f). According to the FeCl2 reference, Fe2+ ions are featured by two shoulders on L2 and L3 edges (indicated by the arrows in Figure 2f) which are absent in the spectrum of ultrathin LDH but present in that of bulk LDH, implying almost pure Fe3+ ions in ultrathin LDH but mixed Fe2+/ Fe3+ ions in bulk LDH. Quantitative analysis shows that ratios of Fe3+:Fe2+ in ultrathin and bulk LDHs are 99.8:0.2 and 85.7:14.3, respectively. The difference is not pronounced enough to explain the relatively large change in Fe K-edge positions (1.38 eV) in Figure 2d; hence, the reason is more likely to be the higher Fe−O covalency in ultrathin LDH.37 This assumption is confirmed by the charge-density wave calculation. As shown in Figure S2, the electrons in ultrathin LDH are more delocalized than those in bulk LDH, indicating the higher Fe−O covalency in ultrathin LDH. It has been reported that the electron delocalization can promote the electron transfer and lower oxidization

1c). Meanwhile, a bulk LDH sample, synthesized by a hydrothermal method (see details in Experimental Methods), also presents a hexagon-like morphology (Figure 1b) but a larger thickness of about 9.5 nm (Figure 1d). By comparing the thickness of ultrathin LDH with the (001) interplanar spacing of α-Ni(OH)2 (Scheme S2), we find that the total thickness of ultrathin LDH accords with a height of 1/3 unit-cell of αNi(OH)2, implying that the ultrathin LDH consists of two unit layers and hence has 100% of its metal sites exposed to the surface. In the X-ray diffraction (XRD) pattern of ultrathin LDH (Figure 1e), only two peaks are observed at ∼11.5° and ∼23°, corresponding to the (003) and (006) planes of αNi(OH)2. Hence, the ultrathin LDH exposes its (001) basal surface. In contrast, bulk LDH shows weak peaks at higher diffraction angles, corresponding to the (101), (012), (015), (018), (110), and (113) planes. The Raman spectra of both samples (Figure 1f) include three peaks at 295, 430, and 530 cm−1, which can be assigned as E-type, M−O(H), and M−O vibrations, respectively.27−36 The I530/I430 ratio of ultrathin LDH is higher than that of the bulk one, indicating the number of M−O motifs is higher than that of Ni−OH motifs. X-ray 6028

DOI: 10.1021/acscatal.9b01935 ACS Catal. 2019, 9, 6027−6032

Letter

ACS Catalysis barrier,39,40 thus facilitating the formation of Fe3+ ions in ultrathin LDH. We then compare the electrochemical water oxidation properties of the two samples, with commercial RuO2 as a reference. Ultrathin LDH presents much higher performance than bulk LDH and RuO2 in 1 M KOH (Figure 3a).

that of bulk LDH. Moreover, the long-term durability was investigated by performing a chronopotentiometric test at 10 mA/cm2. The results show that the ultrathin LDH can endure the long-term testing without obvious degradation of the performance (Figure S7). Soft XAS analysis shows that the Ni L-edge and the Fe L-edge of ultrathin LDH remain unchanged after 80 CV cycles (Figure S8), indicating a high reversibility upon the oxidization and reduction process. To understand the chemical origin of the performance variation between ultrathin LDH and its bulk counterpart, we conducted operando hard XAS to evaluate the changes of oxidation states for metal cations at different working potentials. For Ni ions, operando XAS illustrates that the Ni K-edge of ultrathin LDH shows an obvious positive shift as the applied potential increases, while that of bulk LDH changes slightly (Figures 4a and S9). Since the K-edge position

Figure 3. Electrochemical properties of ultrathin and bulk LDHs. (a) Cyclic voltammetry curves at 5 mV/s with 85%iR correction. R represents the resistance of the electrolyte, determined by the EIS test. (b) Tafel plots. (c) EIS profiles. (d) Amount of active Ni sites determined by the Ni reduction peak. (e) TOF plot. (f) Specific activity of bulk and ultrathin LDH with 5% iR correction.

Specifically, the overpotentials of ultrathin LDH are 210 mV @ 10 mA/cm2 and 320 mV @ 500 mA/cm2, much lower than those of bulk LDH (251 mV @ 10 mA/cm2 and 460 mV @ 500 mA/cm2) (Figures 3a and S3). The lower overpotential of the ultrathin LDH could be ascribed to its higher electrochemically active surface area (ECSA) (Figure S4 and Table S2), which increases the contact area between the catalysts and the electrolyte and facilitates the mass transfer. In addition, the Tafel slope of ultrathin LDH is only ∼30 mV/Dec, lower than the corresponding slopes of bulk LDH (46 mV/Dec) and RuO2 (106 mV/Dec) (Figure 3b). In the EIS spectra, ultrathin LDH exhibits the lowest semicircle, indicating a fast charge transfer between the electrode and the electrolyte (Figure 3c). When compared with the latest published data, ultrathin LDH achieves a top performance among the state-of-the-art NiFe LDH materials (Table S1). Next, we calculate TOF to determine the intrinsic OER activity (Figures 3d,e, S5, and S6). Two methods are employed to determine the number of active sites. The first one takes all the metal ions as catalytically active sites (TOFTM). In this case, TOF is usually underestimated because only part of the metal ions can serve as active sites.43 The second method assumes the charged nickel ions as the active sites (TOFint).41 The number of active sites is determined by integrating the reduction peak of Ni2+/Ni3+/4+ (Figure 3d), so as to avoid the influence of the OER current on the anodic peak during the positive scan. The TOFTM of the ultrathin LDH is higher than that of the bulk LDH, and both LDHs are superior to the commercial RuO2 (Figure S6). In contrast, TOFint values of ultrathin and bulk LDHs are similar (∼0.6 s−1), which implies similar OER mechanisms and active sites for the ultrathin and bulk LDH samples. We also calculate the specific activity by normalizing total current density with ECSA (Figure S4 and Table S2). As shown in Figure 3f, the specific current density of ultrathin LDH is still higher than

Figure 4. Operando XAS results of Ni and Fe element in ultrathin LDHs. (a) Ni K-edge, (b) Fe K-edge, (c) M−O bond length, and (d) coordination number evolution with increasing applied potential during the OER process. The error bars were generated on the basis of the EXAFS fitting errors.

represents the oxidation state,37,42 we conclude that Ni ions in ultrathin LDH are oxidized more easily than those in bulk LDH. According to the Ni K-edge data reported in the literature43 and delithiated LiNiO2 data (Figure S10), the Ni2+ cations in ultrathin LDH could be fully oxidized to Ni4+ at the applied potential of 1.4 V vs reversible hydrogen electrode (RHE) (see Figures 4a, S9, and S10), while for bulk LDH, the average oxidation state is less than 3+ at the same potential (Figures 4a, S9, and S10). Operando EXAFS reveals that the Ni−O bonds in ultrathin LDH shrink from 2.06 to 1.89 Å upon the applied potential (Figures 4a, S11 and Tables S3, S4), confirming the conversion of Ni2+ into Ni4+.43−48 As for Fe ions, the operando Fe K-edge spectra of ultrathin and bulk LDHs shift gradually to positive direction as the applied potential increases (Figures 4b and S12). For ultrathin LDH, the doublet peak in the pre-edge region becomes a singlet peak as the applied potential is above 1.4 V vs RHE (Figure S12a); meanwhile, the intensity of the white line gradually decreases, and an obvious shoulder emerges at 7150 eV (Figure S12b), indicating a dramatic increase in Fe 6029

DOI: 10.1021/acscatal.9b01935 ACS Catal. 2019, 9, 6027−6032

Letter

ACS Catalysis Notes

oxidation state. In addition, the operando EXAFS spectrum of ultrathin LDH indicates that Fe−O bonds contract from 1.99 to 1.88 Å (Figures 4c, S13 and Tables S3, S5), suggesting an iron valence state transformation from Fe3+ to Fe4+.43−45,49−51 In comparison, bulk LDH does not show any obvious change even when the applied potential is as high as 1.5 V vs RHE (Figure S12c,d), implying that the Fe cations in bulk LDH are harder to be oxidized. The electrochemical water oxidation reaction on the ultrathin sample starts at ∼1.42 V vs RHE, corresponding to the formation of tetravalent Fe and Ni (Figure 4a−c). These results indicate that tetravalent Ni and Fe are prerequisites for the OER. On the other hand, the coordination numbers (CNs) for Ni−O and Fe−O bonds in the ultrathin sample were the same with those in the bulk material (∼6) prior to the OER test, implying that Ni and Fe ions are in the octahedral coordination. After the catalyst was activated for the OER, the CNs for Ni and Fe decreased to 5 (Figure 4d), indicating that both Ni- and Fe-centered octahedra lose one of the bonded oxygens. The lattice-oxygen-involved OER process is further supported by the pH-dependent OER behavior of the ultrathin LDH (Figure S14), which is known as the characteristic of the decoupled proton/electron transfer and the loss of lattice oxygen.44,52−54 Therefore, we conclude that the OER occurs on the tetravalent Ni and Fe ions following a decoupled proton/electron mechanism. In summary, ultrathin Ni−Fe LDH with 100% exposed metal ions was successfully produced via an alcohol intercalation process and used as an ideal platform for operando synchrotron X-ray analysis to probe the catalytic mechanism without any “bulk” or subsurface interferences. Under low OER potentials, all metal ions in ultrathin LDH can be fully oxidized into tetravalency state, which facilitates the lattice-oxygen-involved OER in a decoupled proton/electron way. Moreover, since the high surface area improves the mass transportation during the OER, eventually, the ultrathin LDH shows high activity toward water oxidation, with overpotentials of 210 mV @10 mA/cm2. Our findings indicate that the complete oxidization of metal ions is a prerequisite for highly active NiFe LDHs, which can be accomplished by ultrathin (single or double unit layer) structures.



The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work at Tianjin university was supported by the Natural Science Foundation of China (Nos. 51871160, 51671141, and 51471115). The work at Virginia Tech was supported by Department of Chemistry startup funds and the Institute for Critical Technology and Applied Science. The Stanford Synchrotron Radiation Lightsource, a Directorate of SLAC National Accelerator Laboratory and an Office of Science User Facility is operated for the U.S. Department of Energy Office of Science by Stanford University. Use of the Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, is supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract No.DE-AC02-76SF00515.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscatal.9b01935.



REFERENCES

(1) Seh, Z. W.; Kibsgaard, J.; Dickens, C. F.; Chorkendorff, I.; Norskov, J. K.; Jaramillo, T. F. Combining Theory and Experiment in Electrocatalysis: Insights into Materials Design. Science 2017, 355, eaad4998. (2) Zheng, Y.; Jiao, Y.; Zhu, Y.; Li, L. H.; Han, Y.; Chen, Y.; Jaroniec, M.; Qiao, S. Z. High Electrocatalytic Hydrogen Evolution Activity of an Anomalous Ruthenium Catalyst. J. Am. Chem. Soc. 2016, 138, 16174−16181. (3) Liu, M.; Pang, Y.; Zhang, B.; De Luna, P.; Voznyy, O.; Xu, J.; Zheng, X.; Dinh, C. T.; Fan, F.; Cao, C.; De Arquer, F. P.; Safaei, T. S.; Mepham, A.; Klinkova, A.; Kumacheva, E.; Filleter, T.; Sinton, D.; Kelley, S. O.; Sargent, E. H. Enhanced Electrocatalytic CO2 Reduction via Field-Induced Reagent Concentration. Nature 2016, 537, 382−386. (4) Jiao, Y.; Zheng, Y.; Chen, P.; Jaroniec, M.; Qiao, S. Z. Molecular Scaffolding Strategy with Synergistic Active Centers to Facilitate Electrocatalytic CO2 Reduction to Hydrocarbon/Alcohol. J. Am. Chem. Soc. 2017, 139, 18093−18100. (5) Wang, S.; Hai, X.; Ding, X.; Chang, K.; Xiang, Y.; Meng, X.; Yang, Z.; Chen, H.; Ye, J. Light-Switchable Oxygen Vacancies in Ultrafine Bi5O7Br Nanotubes for Boosting Solar-Driven Nitrogen Fixation in Pure Water. Adv. Mater. 2017, 29, 1701774. (6) Bao, D.; Zhang, Q.; Meng, F. L.; Zhong, H. X.; Shi, M. M.; Zhang, Y.; Yan, J. M.; Jiang, Q.; Zhang, X. B. Electrochemical Reduction of N2 under Ambient Conditions for Artificial N2 Fixation and Renewable Energy Storage Using N2/NH3 Cycle. Adv. Mater. 2017, 29, 1604799. (7) Chen, S.; Perathoner, S.; Ampelli, C.; Mebrahtu, C.; Su, D.; Centi, G. Electrocatalytic Synthesis of Ammonia at Room Temperature and Atmospheric Pressure from Water and Nitrogen on a Carbon-Nanotube-Based Electrocatalyst. Angew. Chem., Int. Ed. 2017, 56, 2699−2703. (8) Li, S. J.; Bao, D.; Shi, M. M.; Wulan, B. R.; Yan, J. M.; Jiang, Q. Amorphizing of Au Nanoparticles by CeOx−RGO Hybrid Support Towards Highly Efficient Electrocatalyst for N2 Reduction Under Ambient Conditions. Adv. Mater. 2017, 29, 1700001. (9) Man, I. C.; Su, H.-Y.; Calle-Vallejo, F.; Hansen, H. A.; Martínez, J. I.; Inoglu, N. G.; Kitchin, J.; Jaramillo, T. F.; Nørskov, J. K.; Rossmeisl, J. Universality in Oxygen Evolution Electrocatalysis on Oxide Surfaces. ChemCatChem 2011, 3, 1159−1165. (10) McCrory, C. C.; Jung, S.; Peters, J. C.; Jaramillo, T. F. Benchmarking Heterogeneous Electrocatalysts for the Oxygen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 16977−16987. (11) Burke, M. S.; Zou, S.; Enman, L. J.; Kellon, J. E.; Gabor, C. A.; Pledger, E.; Boettcher, S. W. Revised Oxygen Evolution Reaction Activity Trends for First-Row Transition-Metal (Oxy)hydroxides in Alkaline Media. J. Phys. Chem. Lett. 2015, 6, 3737−3742.

Synthesis of ultrathin Ni−Fe LDH, synthesis of bulk Ni−Fe LDH, characterization, electrochemical test, XAS test, DFT calculation, and additional experimental results (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail for X.W.D.: [email protected]. *E-mail for F.L.: [email protected]. *E-mail for D.S.: [email protected]. ORCID

Feng Lin: 0000-0002-3729-3148 Xi-Wen Du: 0000-0002-2811-147X Author Contributions ¶

(C.K., Y.Z.) These authors contributed equally. 6030

DOI: 10.1021/acscatal.9b01935 ACS Catal. 2019, 9, 6027−6032

Letter

ACS Catalysis (12) Lu, X.; Zhao, C. Electrodeposition of Hierarchically Structured Three-Dimensional Nickel−Iron Electrodes for Efficient Oxygen Evolution at High Current Densities. Nat. Commun. 2015, 6, 6616. (13) Hoang, T. T. H.; Gewirth, A. A. High Activity Oxygen Evolution Reaction Catalysts From Additive-Controlled Electrodeposited Ni and NiFe Films. ACS Catal. 2016, 6, 1159−1164. (14) Gong, M.; Li, Y.; Wang, H.; Liang, Y.; Wu, J. Z.; Zhou, J.; Wang, J.; Regier, T.; Wei, F.; Dai, H. An Advanced Ni−Fe Layered Double Hydroxide Electrocatalyst for Water Oxidation. J. Am. Chem. Soc. 2013, 135, 8452−8455. (15) Friebel, D.; Louie, M. W.; Bajdich, M.; Sanwald, K. E.; Cai, Y.; Wise, A. M.; Cheng, M.; Sokaras, D.; Weng, T.; Alonso-Mori, R.; Davis, R. C.; Bargar, J. R.; Nørskov, J. K.; Nilsson, A.; Bell, A. T. Identification of Highly Active Fe Sites in (Ni,Fe)OOH for Electrocatalytic Water Splitting. J. Am. Chem. Soc. 2015, 137 (3), 1305−1313. (16) Merte, L. R.; Behafarid, F.; Miller, D. J.; Friebel, D.; Cho, S.; Mbuga, F.; Sokaras, D.; Alonso-Mori, R.; Weng, T.; Nordlund, D.; Nilsson, A.; Roldan Cuenya, B. Electrochemical Oxidation of SizeSelected Pt Nanoparticles Studied Using in situ High-EnergyResolution X-ray Absorption Spectroscopy. ACS Catal. 2012, 2 (11), 2371−2376. (17) Gorlin, Y.; Chung, C.; Benck, J. D.; Nordlund, D.; Seitz, L.; Weng, T.; Sokaras, D.; Clemens, B. M.; Jaramillo, T. F. Understanding Interactions between Manganese Oxide and Gold that Lead to Enhanced Activity for Electrocatalytic Water Oxidation. J. Am. Chem. Soc. 2014, 136 (13), 4920−4926. (18) Friebel, D.; Bajdich, M.; Yeo, B. S.; Louie, M. W.; Miller, D. J.; Sanchez Casalongue, H.; Mbuga, F.; Weng, T.; Nordlund, D.; Sokaras, D.; Alonso-Mori, R.; Bell, A. T.; Nilsson, A. On the Chemical State of Co Oxide Electrocatalysts During Alkaline Water Splitting. Phys. Chem. Chem. Phys. 2013, 15, 17460−17467. (19) Friebel, D.; Mbuga, F.; Rajasekaran, S.; Miller, D. J.; Ogasawara, H.; Alonso-Mori, R.; Sokaras, D.; Nordlund, D.; Weng, T.; Nilsson, A. Structure, Redox Chemistry, and Interfacial Alloy Formation in Monolayer and Multilayer Cu/Au(111) Model Catalysts for CO2 Electroreduction. J. Phys. Chem. C 2014, 118 (15), 7954−7961. (20) Garcia-Esparza, A. T.; Shinagawa, T.; Ould-Chikh, S.; Qureshi, M.; Peng, X.; Wei, N.; Anjum, D. H.; Clo, A.; Weng, T.; Nordlund, D.; Sokaras, D.; Kubota, J.; Domen, K.; Takanabe, K. An OxygenInsensitive Hydrogen Evolution Catalyst Coated by a MolybdenumBased Layer for Overall Water Splitting. Angew. Chem., Int. Ed. 2017, 56 (21), 5780−5784. (21) Chakthranont, P.; Kibsgaard, J.; Gallo, A.; Park, J.; Mitani, M.; Sokaras, D.; Kroll, T.; Sinclair, R.; Mogensen, M. B.; Jaramillo, T. F. Effects of Gold Substrates on the Intrinsic and Extrinsic Activity of High-Loading Nickel-Based Oxyhydroxide Oxygen Evolution Catalysts. ACS Catal. 2017, 7 (8), 5399−5409. (22) Frydendal, R.; Seitz, L. C.; Sokaras, D.; Weng, T.; Nordlund, D.; Chorkendorff, I.; Stephens, I. E. L.; Jaramillo, T. F. Operando Investigation of Au-MnOx Thin Films with Improved Activity for the Oxygen Evolution Reaction. Electrochim. Acta 2017, 230, 22−28. (23) Song, F.; Hu, X. Exfoliation of Layered Double Hydroxides for Enhanced Oxygen Evolution Catalysis. Nat. Commun. 2014, 5, 4477. (24) Kuroda, Y.; Miyamoto, Y.; Hibino, M.; Yamaguchi, K.; Mizuno, N. Tripodal Ligand-Stabilized Layered Double Hydroxide Nanoparticles with Highly Exchangeable CO32‑. Chem. Mater. 2013, 25, 2291−2296. (25) Jung, S.; McCrory, C. C. L.; Ferrer, I. M.; Peters, J. C.; Jaramillo, T. F. Benchmarking Nanoparticulate Metal Oxide Electrocatalysts for the Alkaline Water Oxidation Reaction. J. Mater. Chem. A 2016, 4, 3068−3076. (26) Zhao, Y.; Zhang, X.; Jia, X.; Waterhouse, G. I. N.; Shi, R.; Zhang, X.; Zhan, F.; Tao, Y.; Wu, L.-Z.; Tung, C.-H.; O’Hare, D.; Zhang, T. Sub-3 nm Ultrafine Monolayer Layered Double Hydroxide Nanosheets for Electrochemical Water Oxidation. Adv. Energy Mater. 2018, 8, 1703585.

(27) Melendres, C. A.; Xu, S. In situ Laser Raman Spectroscopic Study of Anodic Corrosion Films on Nickel and Cobalt. J. Electrochem. Soc. 1984, 131, 2239−2243. (28) Desilvestro, J.; Corrigan, D. A.; Weaver, M. J. Spectroelectrochemistry of Thin Nickel Hydroxide Films on Gold Using Surfaceenhanced Raman Spectroscopy. J. Phys. Chem. 1986, 90, 6408−6411. (29) Desilvestro, J.; Corrigan, D. A.; Weaver, M. J. Characterization of Redox States of Nickel Hydroxide Film Electrodes by in situ Surface Raman Spectroscopy. J. Electrochem. Soc. 1988, 135, 885−892. (30) Johnston, C.; Graves, P. R. In Situ Raman Spectroscopy Study of the Nickel Oxyhydroxide Electrode (NOE) System. Appl. Spectrosc. 1990, 44, 105−115. (31) Oblonsky, L. J.; Devine, T. M. Surface Enhanced Raman Spectra From the Films Formed on Nickel in the Passive and Transpassive Regions. J. Electrochem. Soc. 1995, 142, 3677−3682. (32) Kostecki, R.; McLarnon, F. Electrochemical and in situ Raman Spectroscopic Characterization of Nickel Hydroxide Electrodes I. Pure Nickel Hydroxide. J. Electrochem. Soc. 1997, 144, 485−493. (33) Balasubramanian, M.; Melendres, C. A.; Mini, S. X-ray Absorption Spectroscopy Studies of the Local Atomic and Electronic Structure of Iron Incorporated into Electrodeposited Hydrous Nickel Oxide Films. J. Phys. Chem. B 2000, 104, 4300−4306. (34) Bantignies, J. L.; Deabate, S.; Righi, A.; Rols, S.; Hermet, P.; Sauvajol, J. L.; Henn, F. New Insight into the Vibrational Behavior of Nickel Hydroxide and Oxyhydroxide Using Inelastic Neutron Scattering, Far/Mid-Infrared and Raman Spectroscopies. J. Phys. Chem. C 2008, 112, 2193−2201. (35) Hall, D. S.; Lockwood, D. J.; Poirier, S.; Bock, C.; MacDougall, B. R. Raman and Infrared Spectroscopy of α and β Phases of Thin Nickel Hydroxide Films Electrochemically Formed on Nickel. J. Phys. Chem. A 2012, 116, 6771−6784. (36) Luo, K.; Roberts, M. R.; Hao, R.; Guerrini, N.; Pickup, D. M.; Liu, Y. S.; Edstrom, K.; Guo, J.; Chadwick, A. V.; Duda, L. C.; Bruce, P. G. Charge-Compensation in 3d-Transition-Metal-Oxide Intercalation Cathodes through the Generation of Localized Electron Holes on Oxygen. Nat. Chem. 2016, 8, 684−691. (37) Shulman, G. R.; Yafet, Y.; Eisenberger, P.; Blumberg, W. E. Observations and Interpretation of X-ray Absorption Edges in Iron Compounds and Proteins. Proc. Natl. Acad. Sci. U. S. A. 1976, 73, 1384−1388. (38) Westre, T. E.; Kennepohl, P.; DeWitt, J. G.; Hedman, B.; Hodgson, K. O.; Solomon, E. I. A Multiplet Analysis of Fe K-edge 1s→3d Pre-edge Features of Iron Complexes. J. Am. Chem. Soc. 1997, 119, 6297−6314. (39) Chen, S.; Kang, Z.; Zhang, X.; Xie, J.; Wang, H.; Shao, W.; Zheng, X.; Yan, W.; Pan, B.; Xie, Y. Highly Active Fe Sites in Ultrathin Pyrrhotite Fe7S8 Nanosheets Realizing Efficient Electrocatalytic Oxygen Evolution. ACS Cent. Sci. 2017, 3, 1221−1227. (40) Chen, S.; Kang, Z.; Hu, X.; Zhang, X.; Wang, H.; Xie, J.; Zheng, X.; Yan, W.; Pan, B.; Xie, Y. Delocalized Spin States in 2D Atomic Layers Realizing Enhanced Electrocatalytic Oxygen Evolution. Adv. Mater. 2017, 29, 1701687. (41) Stevens, M. B.; Enman, L. J.; Batchellor, A. S.; Cosby, M. R.; Vise, A. E.; Trang, C. D. M.; Boettcher, S. W. Measurement Techniques for the Study of Thin Film Heterogeneous Water Oxidation Electrocatalysts. Chem. Mater. 2017, 29, 120−140. (42) Dau, H.; Liebisch, P.; Haumann, M. X-ray Absorption Spectroscopy to Analyze Nuclear Geometry and Electronic Structure of Biological Metal CentersPotential and Questions Examined with Special Focus on the Tetra-Nuclear Manganese Complex of Oxygenic Photosynthesis. Anal. Bioanal. Chem. 2003, 376, 562−583. (43) Görlin, M.; Chernev, P.; Ferreira de Araújo, J.; Reier, T.; Dresp, S.; Paul, B.; Krähnert, R.; Dau, H.; Strasser, P. Oxygen Evolution Reaction Dynamics, Faradaic Charge Efficiency, and the Active Metal Redox States of Ni−Fe Oxide Water Splitting Electrocatalysts. J. Am. Chem. Soc. 2016, 138, 5603−5614. (44) Trzesniewski, B. J.; Diaz-Morales, O.; Vermaas, D. A.; Longo, A.; Bras, W.; Koper, M. T.; Smith, W. A. In situ Observation of Active Oxygen Species in Fe-containing Ni-based Oxygen Evolution 6031

DOI: 10.1021/acscatal.9b01935 ACS Catal. 2019, 9, 6027−6032

Letter

ACS Catalysis Catalysts: The Effect of pH on Electrochemical Activity. J. Am. Chem. Soc. 2015, 137, 15112−15121. (45) Gorlin, M.; Ferreira de Araujo, J.; Schmies, H.; Bernsmeier, D.; Dresp, S.; Gliech, M.; Jusys, Z.; Chernev, P.; Kraehnert, R.; Dau, H.; Strasser, P. Tracking Catalyst Redox States and Reaction Dynamics in Ni−Fe Oxyhydroxide Oxygen Evolution Reaction Electrocatalysts: The Role of Catalyst Support and Electrolyte pH. J. Am. Chem. Soc. 2017, 139, 2070−2082. (46) Balasubramanian, M.; Melendres, C. A.; Mini, S. X-ray Absorption Spectroscopy Studies of the Local Atomic and Electronic Structure of Iron Incorporated into Electrodeposited Hydrous Nickel Oxide Films. J. Phys. Chem. B 2000, 104, 4300−4306. (47) Morishita, M.; Ochiai, S.; Kakeya, T.; Ozaki, T.; Kawabe, Y.; Watada, M.; Tanase, S.; Sakai, T. Structural Analysis by Synchrotron XRD and XAFS for Manganese-Substituted α- and β-Type Nickel Hydroxide Electrode. J. Electrochem. Soc. 2008, 155, A936−A944. (48) Currie, D. B.; Levason, W.; Oldroyd, R. D.; Weller, M. T. Synthesis, Spectroscopic and Structural Studies of Alkali Metal− Nickel Periodates MNiIO6(M = Na, K, Rb, Cs or NH4). J. Chem. Soc., Dalton Trans. 1994, 1483−1487. (49) Thoral, S.; Rose, J.; Garnier, J. M.; Van Geen, A.; Refait, M. P.; Traverse, A.; Fonda, E.; Nahon, D.; Bottero, J. Y. XAS Study of Iron and Arsenic Speciation during Fe(II) Oxidation in the Presence of As(III). Environ. Sci. Technol. 2005, 39, 9478. (50) Suzuki, S.; Suzuki, T.; Kimura, M.; Takagi, Y.; Shinoda, K.; Tohji, K.; Waseda, Y. EXAFS Characterization of Ferric Oxyhydroxides. Appl. Surf. Sci. 2001, 169−170, 109. (51) Hodges, J. P.; Short, S.; Jorgensen, J. D.; Xiong, X.; Dabrowski, B.; Mini, S. M.; Kimball, C. W. Evolution of Oxygen-Vacancy Ordered Crystal Structures in the Perovskite Series SrnFenO3n‑1 (n = 2, 4, 8, and ∞), and the Relationship to Electronic and Magnetic Properties. J. Solid State Chem. 2000, 151, 190. (52) Grimaud, A.; Diaz-Morales, O.; Han, B.; Hong, W. T.; Lee, Y. L.; Giordano, L.; Stoerzinger, K. A.; Koper, M. T. M.; Shao-Horn, Y. Activating Lattice Oxygen Redox Reactions in Metal Oxides to Catalyse Oxygen Evolution. Nat. Chem. 2017, 9, 457−465. (53) Koper, M. T. M. Thermodynamic Theory of Multi-Electron Transfer Reactions: Implications for Electrocatalysis. J. Electroanal. Chem. 2011, 660, 254−260. (54) Koper, M. T. M. Theory of Multiple Proton−Electron Transfer Reactions and its Implications for Electrocatalysis. Chemical science 2013, 4, 2710.

6032

DOI: 10.1021/acscatal.9b01935 ACS Catal. 2019, 9, 6027−6032