Vibronic Super-exchange in Double Perovskite Electrocatalyst for

Publication Date (Web): August 22, 2018 ... Vibronic super-exchange of Ni3+-O-Mn3+ in La2NiMnO6 nanoparticle brings the optimal eg electron filling-st...
8 downloads 0 Views 1MB Size
Subscriber access provided by EKU Libraries

Communication

Vibronic Super-exchange in Double Perovskite Electrocatalyst for Efficient Electrocatalytic Oxygen Evolution Yun Tong, Junchi Wu, Pengzuo Chen, Haifeng Liu, Wangsheng Chu, Changzheng Wu, and Yi Xie J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b06108 • Publication Date (Web): 22 Aug 2018 Downloaded from http://pubs.acs.org on August 22, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Vibronic Super-exchange in Double Perovskite Electrocatalyst for Efficient Electrocatalytic Oxygen Evolution Yun Tong,†,§ Junchi Wu,†,§ Pengzuo Chen,†,§ Haifeng Liu,‡ Wangsheng Chu,⊥ Changzheng Wu,†,* and Yi Xie† †

Hefei National Laboratory for Physical Sciences at the Microscale, iChEM (Collaborative Innovation Center of Chemistry for Energy Materials), and CAS Key Laboratory of Mechanical Behavior and Design of Materials, University of Science and Technology of China, Hefei, Anhui 230026, P.R. China. ⊥

National Synchrotron Radiation Laboratory, University of Science and Technology of China, Hefei, Anhui 230026, P.R. China. ‡

Analytical and Testing Center, Southwest University of Science and Technology, Mianyang, Sichuan 621010, P. R. China

Supporting Information Placeholder ABSTRACT: Perovskites are prototype electrocatalyts benefiting from their terrestrial plenty and high stability. Electronic state regulation plays a key role in promising higher electrocatalytic efficiencies. Herein, we highlighted a vibronic super-exchange in double perovskite to synergistically optimize eg electron filling-state and increase the formation of active species on surface of catalysts. Vibronic super3+ 3+ exchange of Ni -O-Mn in La2NiMnO6 nanoparticle brings the optimal eg electron filling-state of Mn and Ni ions towards unity. Moreover, vibronic super-exchange interaction 3+ 3+ of Ni -O-Mn induces strong Jahn-Teller distortion of MnO6 and NiO6 octahedron, elongating metal-O bonds, which helps to form the active species of Mn/Ni hydroxide/oxide on the surface of catalysts. Surprisingly, La2NiMnO6 nanoparticle exhibits superior OER catalytic performance with higher current density and lower Tafel slope than its bulk counterpart. Our finding will be a promising pathway to developed advanced precious-metal free catalysts.

Driven by growing concerns of environmental pollution and petroleum feedstock’s depletion, developing alternative energy storage and conversion systems is crucial for societal 1,2 goal of sustainable energy. Oxygen evolution reaction (OER), as one of important half reaction of water splitting is greatly hampered by a sluggish kinetics process caused by a complex four-electron oxidation course. Therefore, it needs a high overpotential to reach a desirable current density, leading to a significant loss of overall efficiency of water split3,4 ting. At present, state-of-art OER electrocatalysts are pre5,6 cious metals, such as IrO2 or RuO2. However, the costliness and terrestrial scarcity of these precious-metal based cata7-9 lysts severely hamper their wide-scale applications. Thus, it is of prime importance to design cost-effective and highly-

efficient OER electrocatalysts with lower overpotentials and 10,11 Perovskite oxides with the formula of ABO3, Tafel slopes. where A site and B site are commonly rare-earth and transition-metal ions, respectively, have been considered as promising precious-metal-free electrocatalysts due to their unique 12-14 3d electronic structures and structure flexibility. The previous literatures have reported an experimental principle that when the eg electron filling-state of B-site metal towards a unity, the perovskites would exhibit the highest OER cata15,16 lytic activity. For instance, controllable doping and defects were introduced into the framework to optimize the eg electron filling-state, and finally achieved excellent OER catalytic 17 18 performance in SrCo0.95P0.05O3-δ and Ca0.9Yb0.1MnO3-δ. Double perovskites (A2BB′O6), in which B-sites are occupied alternately by different cations B and B′, displays fantastic physical and chemical properties due to enhanced coupling effect by intervening oxygen bridging every B′ and B 19-21 atom pair. Profited from ordered B-site ions arrangement, the physical correlation effects are prominent in double perovskites, generating novel modulation method to their electronic structure and electric transport property, which paves a distinct avenue to modulate OER performance and study 22-24 OER mechanism of perovskite catalysts. On the other hand, more and more theoretically and experimentally investigation has proved that forming active species on the surface of catalysts is another important factor for improving the 25-27 catalytic activity. However, for most of metal-based catalysts, there are numerous strong metal-metal bonds in the crystal structure, which impedes the formation of active spe28 cies (metal hydroxide or oxide). Therefore, finding a strategy to accelerate the formation of active species and simultaneously optimizing the eg electron filling-state of B-site metals is highly desired for enhancing OER performance of double perovskites. Herein, taking double perovskites La2NiMnO6 as a proofof-concept study, we demonstrated a super-exchange effect

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

a

d

b

Page 2 of 6

b

c

e

f

Figure 1 (a) Rietveld refined XRD patterns of LNMO-1, in which the experimental data was marked of red dots, calculated profile of cyan lines, allowed positions of Bragg reflection of green vertical bars and difference curves of blue lines. (b) TEM image of LNMO-1. (c) Distribution diagram for the particle size of LNMO-1 nanoparticle. (d) XPS spectra of Mn 2p for LNMO samples. (e) Mn K-edge XANES spectra of LNMO samples. (f) Ni K-edge XANES spectra of LNMO samples. to redistribute d electrons configuration in B-site cations 4+ 2+ 3+ from Mn -O-Ni static super-exchange effect to Mn -O3+ Ni vibronic super-exchange effect. Via synergistic optimization of eg electron filling-state to unity and accelerative activation on surface of catalysts, La2NiMnO6 nanoparticle with ~33 nm diameter shows superior OER catalytic performance in alkaline medium. This work provides a novel strategy for inspiring the rational design of highly efficient perovskite electrocatalysts. In this work, La2NiMnO6 precursors were synthesized by sol-gel method. And then, La2NiMnO6 precursors were ano o o nealed at 700 C, 900 C and 1300 C to obtain the La2NiMnO6 particles with size ranging from about 30 nm to o bulk (sample annealed at 700 C denoted as LNMO-1; sample o annealed at 900 C denoted as LNMO-2; sample annealed at o 1300 C denoted as LNMO-3). In order to detect the structure and composition of LNMO products, systematic characterizations have been utilized. As shown in Figure 1a and Figure S2, 3, the Rietveld full-profile structure refinement method of XRD pattern has been firstly used to confirm the crystal structure of LNMO products. It can be observed from these figures that experimental profiles of serials of LNMO products fit well with the calculated curves, indicating their pure crystalline phases and no crystalline phase separation occurred in the process of annealed treatment in our samples. Moreover, elemental-mapping images have been provided to analyze the composition of LNMO products. As shown in Figure S4-6, lanthanum (indicated by red color), nickel (indicated by green color), manganese (indicated by purple color) and oxygen (indicated by cyan color) were homogeneous spatial distributed across the whole LNMO products, indicating the high-purity of the three obtained LNMO catalysts. Furthermore, Figure 1b shows the representative TEM image

of LNMO-1 sample, in which the LNMO-1 exhibits nanoparticle morphology with the size distribution in the range of 15 nm to 55 nm. Under careful observation and statistics, we plot the size distribution in Figure 1c and the average size of LNMO-1 is about 33 nm. Moreover, Scherrer equation has also been utilized to calculate the average size of LNMO-1 for more accurate particle size, and the data was listed in Table S1. The calculated average size for LNMO-1 sample is about 31 nm, which agrees well with size distribution counting result. However, with higher annealing temperature, LNMO nanoparticles tend to gather together, and finally transform into chunk morphology of LNMO-3 with the annealed temperao ture of 1300 C (Figure S7, 8). Above experimental results suggest that LNMO products with different size have been successfully synthesized. In order to further study chemical composition and evaluate oxidation state information for the products, X-ray photoelectron spectroscopy (XPS) test has been operated. As shown in Figure S9, the XPS surveys show La, Ni, Mn and O are included in three LNMO samples, which is consistent with the results of elemental mapping analysis. Figure 1d shows the Mn 2p core level spectra of three samples. The XPS peak of Mn 2p3/2 splits into two peaks located at 641.7 eV and 643.9 eV, which well match with the Mn(III) and Mn(IV), 29,30 respectively. More importantly, the Mn(III)/Mn(IV) intensity ratio shows obviously increase with the particle size decreasing, demonstrating the gradual reduction of Mn oxidation state from Mn(IV) to Mn(III). However, since La XPS peaks often strongly overlap with Ni XPS, thus it is not reliable to directly deduce valence state changes of Ni via XPS 31,32 analysis in LNMO. Therefore, in order to gain direct and in-depth investigation of the oxidation state variation for LNMO samples, X-ray absorption near edge structure

ACS Paragon Plus Environment

Page 3 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

a

a

b

c

b

d

c Figure 3 (a) OER polarization curves of LNMO samples. (b) The corresponding Tafel slopes of LNMO samples. (c) Direct comparison of the activity (histogram) and Tafel slope (solid line). (d) Chronoamperometric response of LNMO-1 product. 3+

Figure 2 (a) The temperature dependence of magnetization (M-T) curves under ZFC. (b) The magnetization ratio of Tc1/Tc2 for LNMO samples with different particle size. (c) The schematic for the deformation of MnO6 and NiO6 octahedron. (XANES) of Mn and Ni K-edge has further been performed. As shown in Figure 1e, the Mn K-edge XANES spectra of three LNMO products show similar spectral feature, indicating the whole framework of LNMO maintains intactly after the size regulation. Remarkably, the amplifying rising edge of XANES spectra in the inset of Figure 1e shows an energy shift to lower energy after nanocrystallization, giving further evidence for the decrease of manganese oxidation state as the 33 particle size reduces. Moreover, as shown in Figure 1f and Figure S10, Ni K-edge XANES spectra show a slight shift toward higher energy and narrowed white-line peaks with the size decrease, which further suggests the oxidation state of Ni increases when the LNMO is nanocystallized, indicating the increment of Ni(III) in nanosized LNMO-1. The conversion in valence state of Mn and Ni can effectively modulate d electron arrangement and eg electron number. Magnetism measurement offers an advantaged platform to detect d electron arrangement in our LNMO samples. Zerofield cooling temperature dependent susceptibility curves (M-T curves) of different LNMO samples are utilized to inspect the variation of super-exchange interaction, as shown in Figure 2a. It can be seen that all M-T curves exhibit two phase transition at ~280 K (marked as Tc1) and ~150 K (marked as Tc2). Previous literatures analyzed these two phase transition in detail and pointed out the Tc1 phase orig2+ inates from static super-exchange interaction of Ni -O4+ Mn , while the Tc2 phase comes from vibronic super3+ 3+ 34 exchange interaction of Ni -O-Mn . Of note, with the increase of particle size, the intensity of high Tc1 phase increases while the intensity of low Tc2 phase becomes weaker, 3+ 3+ indicating the suppression of Ni -O-Mn vibronic super35 exchange interaction. As shown in Figure 2b, the susceptibility ratio of Tc1 phase to Tc2 phase raises as the particle size

3+

grow up, suggesting the suppression of Mn and Ni species as confirmed by XPS analysis. Moreover, as schemed in Fig3+ 3+ ure 2c, the spin states of Mn and Ni are high-spin and 3+ low-spin, respectively. High-spin Mn shows an electron 3+ 3 1 configuration of t2g eg and low-spin Ni shows an electron 6 1 configuration of t2g eg ,which induces the deformation of locally cooperative MnO6 and NiO6 octahedron by strong Jahn-Teller effect with elongating metal-O bonds. The elongating metal-O bonds may help to the surface reconstruction for the active species. Both transition metals own an optimal 1 eg electron configuration and therefore the nanosized LNMO is expected to be a high-performance OER catalyst according to the principle between OER electrocatalytic activity and d electron configuration in perovskite-type cata36 lysts. In order to evaluate OER catalytic performance, linearscan voltammetry (LSV) experiments were conducted. Figure 3a shows the IR-corrected LSV curves for all synthesized LNMO catalysts. LNMO-1 product with smallest particle size exhibits an onset potential of 1.54 V vs RHE, which was significantly lower than that of LNMO-2 and LNMO-3. Moreover, at a specific overpotential of 500 mV, the current density 2 of LNMO-1 reaches 101.8 mA/cm , which is much higher than 2 2 that of 30.62 mA/cm for LNMO-2 and 14.19 mA/cm for LNMO-3. Besides, LNMO-1 demonstrated a superior chemical activity compared with benchmark catalyst IrO2 as shown in Figure S11. Tafel slopes are plotted to give further insights into the OER mechanism on the LNMO catalysts. As shown in Figure 3b, Tafel slopes are 58 mV/dec, 73 mV/dec and 91 mV/dec for LNMO-1, LNMO-2 and LNMO-3, respectively, indicating LNMO-1 exhibits the fastest reaction kinetics during the OER process. Moreover, the value of 58 mV/dec for LNMO-1 is also superior to that of most-reported perovskites: i.e., the perovskite BaCo0.5Fe0.4Sn0.1O3-δ catalyst (76 37 38 mV/dec), SrNb0.1Co0.7Fe0.2O3-δ nanorod (61 mV/dec), 39 LaNiO3-δ catalyst (95 mV/dec) and nanosized LaCoO3 (69 40 mV/dec). Figure 3c gives the direct comparison of OER catalytic activity of three LNMO samples with the overpoten2 tial at the current density of 10 mA/cm and Tafel slopes,

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

which clearly shows the greatly enhancement of OER catalytic performance by particle size regulation.

a

test. As shown in Figure 4c, active phase could be hardly seen on pristine LNMO-1 catalyst. After 50 CV tests, a thin active

b

c

d

d d

Page 4 of 6

e

f

Figure 4 (a) OER polarization curves of LNMO-1 sample with CV cycles increasing. (b) The change of current density and overpotential with the increase of CV cycles. (c) HRTEM images of pristine LNMO-1 sample (without activated) and HRTEM images after (d) 50, (e) 100, (f) 500 cycles. Moreover, electrochemical impedance spectroscopy (EIS) experiments showed a dramatically lower interfacial charge transfer resistance on the LNMO-1 catalyst comparing to the other two samples (Figure S12), which indicated much promoted OER kinetics of the LNMO-1 catalyst and accounted for its enhanced performance in the electrochemical catalysis process. Besides, electrochemical double-layer capacitance (Cdl) was introduced to evaluate the active surface area caused by size decreasing. As shown in Figure S13 - S15, Cdl of −2 LNMO-1 is confirmed to be 51.5 mF cm , which is much −2 higher than that of LNMO-2 (Cdl = 41.7 mF cm ) and LNMO−2 3 (Cdl = 32.4 mF cm ). Since Cdl is proportional to the active surface area of electrocatalysts, the results demonstrate that LNMO-1 is more effective in enlarging the catalytically active surface area in comparison to other two products due to the decreasing size; thus, better exposure and enhanced utilization of electroactive sites (e.g., Mn and Ni species). Stability is another important factor to evaluate an electrocatalyst. As shown in Figure 3d, LNMO-1 exhibits excellent stability with negligible recession of current density after OER testing for a long time. Notably, the apparent vibronic super-exchange interaction 3+ 3+ between high-spin Mn and low-spin Ni also can facilitate the formation of active species, i.g., nickel oxide/hydroxide and manganese oxide/hydroxide, on the surface of LNMO-1 catalyst. As shown in Figure 4a, the OER performance of LNMO-1 gets enhancement during the OER process. After 200 cyclic voltammetry (CV) tests, the OER performance tends to be stabilized. Figure 4b gives the detailed observation of OER performance depends on the testing cycle number. The change of current density at 1.77 V and the overpo2 tential at the current density of 10 mA/cm clearly show the enhancement of OER performance during the continuous CV test. In order to detect the active species of OER catalysis, we performed a list of ex-situ HRTEM tests on the surface of LNMO-1 catalyst before (marked as pristine) and after CV

phase with a thickness of about 1 nm occurs on the surface of the catalyst. When prolonging the CV tests, the thickness of electroactive species further increases and gradually reaches a steady state (Figure 4d-f). However, as shown in Figure S1617, the active phase on the surface of LNMO-2 and LNMO-3 catalysts after 500 CV tests are obviously thinner than that of 3+ 3+ LNMO-1, demonstrating the active role of Mn and Ni for forming active species during OER process. The above results 3+ 3+ reveal that the appearance of Mn and Ni could facilitate the formation of active phase on the surface of LNMO sample, for further promoting the catalytic activity. In conclusion, we highlighted a vibronic super-exchange effect in perovskite system to optimize eg electron fillingstate and simultaneously promote active species formation on the surface of catalysts via size regulation. As the particle size of La2NiMnO6 catalysts decreases from bulk to about 33 nm, the super-exchange interaction in La2NiMnO6 transfers 2+ 4+ 3+ from static Ni -O-Mn to vibronic super-exchange of Ni 3+ 1 O-Mn with the eg electron filling-state of both Mn and Ni ions. Moreover, the new super-exchange model with trivalent Mn/Ni ions facilitates the formation of Mn/Ni oxide/hydroxide active species on the surface of La2NiMnO6 nanoparticles. Benefiting from the optimization of eg electron filling-state and accelerating the formation of active phase, the 33-nm La2NiMnO6 nanoparticle shows the best superior OER catalytic performance in alkaline medium. We anticipate that our work could provide new insights for understanding the active phase formation for OER and open a new avenue to design new high performance electrocatalysts.

ASSOCIATED CONTENT Supporting Information Experimental sections, electrochemical test, and other additional figures. This material is available free of charge via the Internet at http://pubs.acs.org.

ACS Paragon Plus Environment

Page 5 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

AUTHOR INFORMATION Corresponding Author [email protected]

Author Contributions §These authors contributed equally to this work.

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT This work was financially supported by the National Basic Research Program of China (2015CB932302), Natural Science Foundation of China (No. 91745113, 11621063, 51502249), National Program for support of Top-notch Yong Professionals, the Fundamental Research Funds for the Central Universities (no. WK2060190084). We also appreciate the support from the Major/Innovative Program of Development Foundation of Hefei Center for physical Science and Technology. This work was partially carried out at the USTC Center for Micro and Nanoscale Research and Fabrication.

REFERENCES (1) (2)

(3) (4) (5)

(6) (7) (8)

(9) (10)

(11) (12) (13) (14) (15) (16)

(17) (18) (19)

Anantharaj, S.; Ede, S. R.; Sakthikumar, K.; Karthick, K.; Mishra, S.; Kundu, S., ACS Catal. 2016, 6 (12), 8069-8097. Gong, M.; Li, Y.; Wang, H.; Liang, Y.; Wu, J. Z.; Zhou, J.; Wang, J.; Regier, T.; Wei, F.; Dai, H., J. Am. Chem. Soc. 2013, 135 (23), 84528455. Kanan, M. W.; Nocera, D. G., Science 2008, 321 (5892), 1072-1075. Trotochaud, L.; Ranney, J. K.; Williams, K. N.; Boettcher, S. W., J. Am. Chem. Soc. 2012, 134 (41), 17253-17261. Reier, T.; Pawolek, Z.; Cherevko, S.; Bruns, M.; Jones, T.; Teschner, D.; Selve, S. r.; Bergmann, A.; Nong, H. N.; Schlögl, R., J. Am. Chem. Soc. 2015, 137 (40), 13031-13040. Tong, Y.; Chen, P.; Zhou, T.; Xu, K.; Chu, W.; Wu, C.; Xie, Y., Angew. Chem. Int. Ed. 2017, 129 (25), 7227-7231. McCrory, C. C.; Jung, S.; Peters, J. C.; Jaramillo, T. F., J. Am. Chem. Soc. 2013, 135 (45), 16977-16987. Xiao, Z.; Wang, Y.; Huang, Y.-C.; Wei, Z.; Dong, C.-L.; Ma, J.; Shen, S.; Li, Y.; Wang, S., Energy Environ. Sci. 2017, 10 (12), 25632569. Minguzzi, A.; Fan, F.-R. F.; Vertova, A.; Rondinini, S.; Bard, A. J., Chem. Sci. 2012, 3 (1), 217-229. Dou, S.; Dong, C. L.; Hu, Z.; Huang, Y. C.; Chen, J. l.; Tao, L.; Yan, D.; Chen, D.; Shen, S.; Chou, S., Adv. Funct. Mater. 2017, 27 (36), 1702546. Wang, Y.; Xie, C.; Zhang, Z.; Liu, D.; Chen, R.; Wang, S., In Situ Exfoliated, Adv. Funct. Mater 2018, 28 (4), 1703363. Royer, S.; Duprez, D.; Can, F.; Courtois, X.; Batiot-Dupeyrat, C.; Laassiri, S.; Alamdari, H., Chem. Rev. 2014, 114 (20), 10292-10368. Hwang, J.; Rao, R. R.; Giordano, L.; Katayama, Y.; Yu, Y.; ShaoHorn, Y., Science 2017, 358 (6364), 751-756. Tong, Y.; Guo, Y.; Chen, P.; Liu, H.; Zhang, M.; Zhang, L.; Yan, W.; Chu, W.; Wu, C.; Xie, Y., Chem 2017, 3 (5), 812-821. Suntivich, J.; May, K. J.; Gasteiger, H. A.; Goodenough, J. B.; Shao-Horn, Y., Science 2011, 334 (6061), 1383-1385. Grimaud, A.; May, K. J.; Carlton, C. E.; Lee, Y.-L.; Risch, M.; Hong, W. T.; Zhou, J.; Shao-Horn, Y., Nat. Commun. 2013, 4, 2439. Zhu, Y.; Zhou, W.; Sunarso, J.; Zhong, Y.; Shao, Z., Adv. Funct. Mater. 2016, 26 (32), 5862-5872. Guo, Y.; Tong, Y.; Chen, P.; Xu, K.; Zhao, J.; Lin, Y.; Chu, W.; Peng, Z.; Wu, C.; Xie, Y., Adv. Mater. 2015, 27 (39), 5989-5994. Vasala, S.; Karppinen, M., Prog. Solid State Chem. 2015, 43 (1-2), 1-36.

(20) Kobayashi, K.-I.; Kimura, T.; Sawada, H.; Terakura, K.; Tokura, Y., Nature 1998, 395 (6703), 677. (21) García-Hernández, M.; Martínez, J.; Martínez-Lope, M.; Casais, M.; Alonso, J., Phys. Rev. Lett. 2001, 86 (11), 2443. (22) Tao, S.; Irvine, J. T., Nat. Mater. 2003, 2 (5), 320. (23) Sun, H.; Chen, G.; Sunarso, J.; Dai, J.; Zhou, W.; Shao, Z., ACS Appl. Mater. Interfaces 2018, 10, 16939−16942 (24) Zhao, B.; Zhang, L.; Zhen, D.; Yoo, S.; Ding, Y.; Chen, D.; Chen, Y.; Zhang, Q.; Doyle, B.; Xiong, X., Nat. Commun. 2017, 8, 14586. (25) Binninger, T.; Mohamed, R.; Waltar, K.; Fabbri, E.; Levecque, P.; Kötz, R.; Schmidt, T. J., Sci. Rep. 2015, 5, 12167. (26) Fabbri, E.; Nachtegaal, M.; Binninger, T.; Cheng, X.; Kim, B.-J.; Durst, J.; Bozza, F.; Graule, T.; Schäublin, R.; Wiles, L., Nat. Mater. 2017, 16 (9), 925. (27) Liang, H.; Gandi, A. N.; Xia, C.; Hedhili, M. N.; Anjum, D. H.; Schwingenschlögl, U.; Alshareef, H. N., ACS Energy Lett. 2017, 2 (5), 1035-1042. (28) Chen, P.; Xu, K.; Fang, Z.; Tong, Y.; Wu, J.; Lu, X.; Peng, X.; Ding, H.; Wu, C.; Xie, Y. Angew. Chem. Int. ed. 2015, 127, 14923 (29) Toupin, M.; Brousse, T.; Bélanger, D., Chem. Mater. 2002, 14 (9), 3946-3952. (30) Ramírez, A.; Hillebrand, P.; Stellmach, D.; May, M. M.; Bogdanoff, P.; Fiechter, S., J. Phys. Chem. C 2014, 118 (26), 1407314081. (31) Qiao, L.; Bi, X., EPL (Europhys. Lett.) 2011, 93 (5), 57002. (32) Abbate, M.; Zampieri, G.; Prado, F.; Caneiro, A.; GonzalezCalbet, J.; Vallet-Regi, M.,. Phys. Rev. B 2002, 65 (15), 155101. (33) Peng, X.; Guo, Y.; Yin, Q.; Wu, J.; Zhao, J.; Wang, C.; Tao, S.; Chu, W.; Wu, C.; Xie, Y., J. Am. Chem. Soc. 2017, 139 (14), 5242-5248. (34) Goodenough, J. B.; Wold, A.; Arnott, R.; Menyuk, N., Phys. Rev. 1961, 124 (2), 373. (35) Zhao, S.; Shi, L.; Zhou, S.; Zhao, J.; Yang, H.; Guo, Y., J. Appl. Phys. 2009, 106 (12), 123901. (36) Hong, W. T.; Risch, M.; Stoerzinger, K. A.; Grimaud, A.; Suntivich, J.; Shao-Horn, Y., Energy Environ. Sci. 2015, 8 (5), 1404-1427. (37) Zhu, Y.; Zhou, W.; Zhong, Y.; Bu, Y.; Chen, X.; Zhong, Q.; Liu, M.; Shao, Z., Adv. Energy Mater. 2017, 7 (8). (38) Zhou, W.; Sunarso, J., J Phys. Chem. Lett 2013, 4 (17), 2982-2988. (39) Xu, X.; Su, C.; Zhou, W.; Zhu, Y.; Chen, Y.; Shao, Z., Adv. Sci. 2016, 3 (2). 1500187 (40) Zhou, S.; Miao, X.; Zhao, X.; Ma, C.; Qiu, Y.; Hu, Z.; Zhao, J.; Shi, L.; Zeng, J., Nat. Commun. 2016, 7, 11510.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table of Contents

ACS Paragon Plus Environment

Page 6 of 6