Next Article in Journal
Generation of Orbital Angular Momentum Modes Using Fiber Systems
Next Article in Special Issue
A New Model for Constant Fuel Utilization and Constant Fuel Flow in Fuel Cells
Previous Article in Journal
Research on the Influence of Backlash on Mesh Stiffness and the Nonlinear Dynamics of Spur Gears
Previous Article in Special Issue
A Theoretical Model for the Triple Phase Boundary of Solid Oxide Fuel Cell Electrospun Electrodes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Controlling the Oxygen Electrocatalysis on Perovskite and Layered Oxide Thin Films for Solid Oxide Fuel Cell Cathodes

1
Department of Mechanical Engineering, University of South Carolina, Columbia, SC 29208, USA
2
Energy Lab, Samsung Advanced Institute of Technology, Suwon, Gyeonggi-do 16678, Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2019, 9(5), 1030; https://doi.org/10.3390/app9051030
Submission received: 13 February 2019 / Revised: 5 March 2019 / Accepted: 6 March 2019 / Published: 12 March 2019
(This article belongs to the Special Issue Progress in Solid-Oxide Fuel Cell Technology)

Abstract

:
Achieving the fast oxygen reduction reaction (ORR) kinetics at the cathode of solid oxide fuel cells (SOFCs) is indispensable to enhance the efficiency of SOFCs at intermediate temperatures. Mixed ionic and electronic conducting (MIEC) oxides such as ABO3 perovskites and Ruddlesden-Popper (RP) oxides (A2BO4) have been widely used as promising cathode materials owing to their attractive physicochemical properties. In particular, oxides in forms of thin films and heterostructures have enabled significant enhancement in the ORR activity. Therefore, we aim to give a comprehensive overview on the recent development of thin film cathodes of SOFCs. We discuss important advances in ABO3 and RP oxide thin film cathodes for SOFCs. Our attention is also paid to the influence of oxide heterostructure interfaces on the ORR activity of SOFC cathodes.

1. Introduction

Solid oxide fuel cells (SOFCs) have shown great promise to contribute to the ultimate aim of environmentally friendly, efficient energy production and conversion [1,2,3,4]. However, overcoming the slow kinetics of the oxygen reduction reaction (ORR) at the cathode is essential to fulfill the increasing demand for developing intermediate temperature SOFCs [5]. At high temperatures such as 1000 °C, La1−xSrxMnO3 (LSM113) is commonly used as a cathode for SOFCs owing to their good stability and high electronic conductivity, but simultaneously has low ionic conductivity and a narrow triple-phase boundary (TPB) region for the cathode reaction [1]. Therefore, new cathode materials with faster ORR kinetics are required to facilitate the development of intermediate temperature SOFCs.
Cubic perovskite based mixed ionic and electronic conductors (MIECs) such as ABO3 oxides [6,7,8,9] and layered perovskite based MIECs such as Ruddlesden-Popper (RP) oxides [10,11,12,13,14] have been extensively studied to promote oxygen electrocatalysis at intermediate temperatures such as 600 °C. Compared to LSM113, these MIECs possess considerably higher oxygen surface exchange kinetics, which enables to extend the electrochemically active region to the entire oxide surface from the electrode/electrolyte interface. In addition, they also have higher ionic conductivity than LSM113, and thus are promising cathode materials for intermediate temperature SOFCs. The oxygen diffusion and surface exchange coefficients for ABO3 and RP oxides are summarized in Table 1.
In ABO3 and RP oxides, the oxygen ion diffusion and oxygen surface exchange kinetics can be significantly influenced by cation substitution as substitutions on the A or B sites of MIECs are known to affect the oxygen stoichiometry. For example, substituting Sr for lanthanum on the A-site of La1−xSrxCoO3−δ (LSC113) results in an increase in oxygen vacancies, enhancing the oxygen transport and surface exchange kinetics [9,15,16,17,18,19]. In addition, doping with different-size cations in ABO3 perovskites can also facilitate oxygen migration owing to a reduction in the ion migration energy [20]. The effect of Sr substitution on the activation energy for oxygen ion diffusion and oxygen surface exchange kinetics in LSC113, reviewed by Berenov et al. [21], is that the activation energies for oxygen diffusion and surface exchange kinetics decrease with increasing Sr content owing to the reduced vacancy formation energy. Cater et al. [22] compared the oxygen diffusivity between Sr-doped cobaltites and Sr-doped manganites, concluding that B-site cations with a lower valence state increase in oxygen diffusivity. In the case of RP oxides, however, the effect of cation substitutions on the ORR kinetics is poorly understood due to difficulties in synthesizing RP oxides with a wide range of cation substitutions.
Table 1. Oxygen self-diffusion (D*) and surface exchange coefficients (k*) for solid oxide fuel cells (SOFC) cathode materials.
Table 1. Oxygen self-diffusion (D*) and surface exchange coefficients (k*) for solid oxide fuel cells (SOFC) cathode materials.
MaterialD*/cm2∙s−1 @ 600 °Ck*/cm∙s−1 @ 600 °CReference
La0.8Sr0.2MnO3−δ10−171.12 × 10−10[8]
La0.5Sr0.5MnO3−δ4.22 × 10−1810−10[8]
La0.8Sr0.2CoO3−δ1.69 × 10−131.38 × 10−8[8]
La0.6Sr0.4CoO3−δ2.86 × 10−99.09 × 10−8[21]
La0.5Sr0.5CoO3−δ2.51 × 10−103.29 × 10−7[8]
La0.6Sr0.4Co0.2Fe0.8O3−δ5.83 × 10−102.35 × 10−7[23]
Ba0.5Sr0.5Co0.8Fe0.2O3−δ1.28 × 10−75.21 × 10−5[24]
La2NiO4+δ1.35 × 10−81.02 × 10−6[13]
La1.9Sr0.1NiO4+δ4.66 × 10−93.56 × 10−8[13]
La1.8Sr0.2NiO4+δ2.02 × 10−103.51 × 10−9[13]
La2CuO4+δ2.76 × 10−95.79 × 10−7[11]
La2Ni0.9Co0.1O4+δ1.07 × 10−86.52 × 10−7[14]
La2Ni0.9Co0.1O4+δ1.34 × 10−87.89 × 10−7[14]
La2CoO4+δ2.98 × 10−83.91 × 10−6[25]
All materials are polycrystalline.
Recently, a remarkable enhancement of the ORR kinetics has been achieved by using epitaxial oxide thin films, which have yielded many intriguing physical and chemical properties that cannot be achieved in bulk counterparts. Epitaxial thin films can also allow the understanding of the fundamental physicochemical properties of oxide materials, and thus have been used as model systems to develop design principles for enhancing the ORR kinetics [26,27,28]. In particular, the use of oxide thin films has the following advantages for manipulating and evaluating electrocatalytic properties [29]. First, epitaxial thin films can be grown with different orientations, which can allow evaluating anisotropic behaviors (appropriate selection of the single crystal substrates and optimization of the deposition conditions). Second, the termination plane of each surface structure can be controlled using epitaxial thin films, and thus the different surface exchange kinetics can be elucidated. Third, dense thin films can allow studying oxygen transport kinetics with no influence of non-kinetic issues, such as particle morphology and connectivity, and porosity and tortuosity typical in porous electrodes. Lastly, the influence of the strain induced by the substrate on the electrocatalytic properties can be investigated. Furthermore, the use of oxide multilayer heterostructures has opened attractive opportunities in the design of new cathode material systems as oxide interfaces can drastically facilitate the ORR kinetics. So far, oxide heterostructure interfaces have been utilized to achieve unique physical and chemical properties such as phase transition [30], electronic reconstruction [31], magnetic domain rearrangement [32], and anomalous oxygen electrocatalysis [33], which are unattainable in the bulk oxides.
This review aims to provide a synopsis of the major developments and achievements in thin film MIEC cathodes of SOFCs, which can help facilitate the development of intermediate temperature SOFCs. We first discuss the ORR mechanisms on the cathode of SOFCs. We then cover in detail recent advances in ABO3 and RP oxide thin film cathodes focusing on cobaltites, ferrites, nickelates, and cuprates. The effect of oxide heterostructure interfaces on the ORR kinetics and electrochemical stability is also overviewed. Finally, we present conclusions, perspectives and future developments. As this review deals with mostly epitaxial thin films, we do not focus on the kinetics of oxygen transport. Readers can be referred to relevant comprehensive reviews of oxygen transport in ABO3 and RP oxide [12,34,35].

2. Oxygen Reduction in SOFCs

2.1. Oxygen Reduction Reaction Steps and Pathways in SOFC Cathodes

Among the entity of charge transport and transfer processes in an operating SOFC, the activation loss η associated with the ORR at the cathode is a main contribution to the overall voltage drop [36]. In MIEC cathodes, ABO3 oxides are generally known as oxygen-deficient perovskites because oxygen vacancies are their dominant anion defect while RP oxides can be both oxygen-deficient and oxygen-excess, depending upon their majority oxygen defects. In the case of oxygen-deficient RP oxides, oxygen nonstoichiometry (δ) arises from oxygen vacancies, whereas oxygen interstitials result in oxygen hyperstochiometric RP oxides. Therefore, two ORR reactions may occur in MIEC cathodes depending on majority oxygen defects [13,37,38]. The overall ORR kinetics for oxygen-excess RP oxides can be expressed as:
O2(g) + 4e → 2Oi
where e is an electron and Oi″ is an oxygen interstitial. Unlike oxygen-excess RP oxides, the following ORR reaction can occur in ABO3 and oxygen deficient RP oxides.
O2(g) + 4e + 2Vo∙∙ → 2Oox
where Vo∙∙ is an oxygen vacancy and Oox represents an oxygen ion in the cathode or electrolyte lattice.
However, ORR reaction is composed of numerous elementary reaction steps. These steps can be surface or bulk related reactions such as adsorption, dissociation, electronation, and diffusion (Figure 1). As shown in Figure 1, pure electronic conductors such as Pt and LSM113 have the surface pathway for the possible ORR reaction steps, which are adsorption and dissociation of molecular oxygen on the cathode surface, surface diffusion of adsorbed oxygen atoms to the TPB, formation of oxide ions by electron transfer and incorporation into the electrolyte. Previous studies [39,40,41,42,43,44] identified using polycrystalline LSM113 samples that the ORR occurs only at the one-dimensional TPB, where cathode, electrolyte, and gas phase are in contact in pure electronic conductors.
In contrast, the surface pathway is complemented by a bulk pathway in MIEC cathodes, where adsorbed oxygen species are reduced at the cathode surface and transported to the electrolyte by solid-state diffusion through the cathode interior regardless of the dominant oxygen defects. Consequently, the active area for the ORR on MIEC cathodes spans over the entire cathode surface. The overall rate of the multitude of reaction steps in MIEC cathodes is determined by a single rate determining step. Then, all remaining steps can be considered to be close to equilibrium so that the voltage loss associated with the rate-determining step constitutes the cathodic activation loss η.

2.2. Oxygen Diffusion and Surface Exchange Kinetics in MIEC Cathodes

The total electrical conductivity σtot of perovskite oxides can be expressed as:
σtot = σel + σion
where σel and σion are electronic conductivity and ionic conductivity, respectively. Depending on the relative amount of σel and σion, perovskite oxides are classified into pure electronic conductors or MIECs.
Oxide cathode materials are typically extrinsic semiconductors, where σel results from the movement of holes or excited electrons through the surrounding lattice. In addition, MIEC cathodes conduct O2- ions via oxygen vacancies or interstitials at a significant rate. The mobility of O2− ions in MIEC cathodes is correlated with the diffusion coefficient (D*) for O2−. In addition, the oxygen surface exchange coefficient (k*) is also a key parameter for determining the rate of oxygen transport through a MIEC cathode material. Therefore, D* and k* play a determining role in the ORR kinetics on MIEC cathodes.
In the case of thin film MIEC cathodes, the ORR kinetics can be limited by either oxygen diffusion or oxygen surface exchange kinetics depending on the critical film thickness, tcrit, which can be defined as D*/k* [45]. The ORR kinetics of thick film cathodes are mainly governed by the bulk transport oxygen whereas in thin film cathodes the ORR kinetics can be governed by the surface exchange of oxygen. In general, tcrit is in the range of 100 μm for many perovskite oxides [46]. Therefore, the ORR kinetics of thin film MIEC cathodes can be enhanced by an increase in the oxygen surface exchange kinetics.
Electrochemical impedance spectroscopy (EIS) measurements are commonly used to analyze the kinetics of oxygen surface exchange on thin film cathodes [6,26,28,47]. Typical EIS tests are conducted in the frequency range from ~10 MHz to ~100 μHz using an AC voltage amplitude with zero DC bias in a wide range of temperatures and oxygen partial pressures. Without dc bias, the Nyquist plot represents three well-separated features analyzed by an equivalent circuit [48], which are denoted as high frequency (HF), medium frequency (MF) and low frequency (LF) feature. The HF feature (104–105 Hz) is associated with the transport of O2− ions through an electrolyte, while the MF features (103–104 Hz) is associated with the ionic transfer resistance at the electrode/electrolyte interface. In addition, the magnitude of the capacitance at the MF feature is relatively small (~10−6 F) compared to the LF feature (~10−3 F), which is the main resistive process representing the surface exchange reaction [6,49,50,51,52]. In the case of thin film cathodes, the magnitude of capacitance is due to the oxygen content change in the films, and therefore the oxygen surface reaction on thin film electrodes generally corresponds with the LF feature. The electrical oxygen surface exchange coefficient (kq) can be determined using the expression [53,54],
kq = R∙T/4F2RORR∙Aelectrode∙co
where R is the universal gas constant (8.314 J·mol−1·K−1), T is the absolute temperature, F is the Faraday’s constant (96,500 C·mol−1), RORR is the resistance obtained from the LF feature, Aelectrode is the area of the electrode, and co is the lattice oxygen concentration in thin film cathodes.

3. Thin Film Cathodes

As discussed earlier, the use of epitaxial thin films offers many advantages for controlling the ORR kinetics in SOFC cathodes. Strain engineering is one of the most effective ways to tailor the ORR kinetics of perovskite and layered oxide thin films. Many studies have reported that epitaxial strain strongly influences bulk and surface properties of oxide materials including the electronic structure [55], oxygen transport [56,57,58,59], oxygen defect formation [60], and oxygen surface exchange kinetics [61,62], ultimately enhancing the high temperature oxygen electrocatalysis in ABO3 and RP oxides [63,64]. In addition, the viability of controlling the crystallographic orientation of epitaxial thin films has enabled investigation of the anisotropic nature of oxygen transport and kinetic properties in RP oxides [61,65]. Furthermore, oxide interface engineering has brought a new concept to design highly active and stable cathode materials for intermediate SOFCs [66,67,68]. The growth of high quality epitaxial thin films is made possible using several deposition methods such as pulsed laser deposition (PLD), molecular beam epitaxy (MBE), atomic layer deposition (ALD), and sputtering. However, it is not the aim of this review to describe in detail those techniques; an extended review on the different techniques can be found elsewhere [69,70,71,72], In the following subsections, we focus on how these factors have brought advances in the oxygen surface exchange kinetics in ABO3 and RP oxide thin films.

3.1. ABO3 Oxide Thin Films

ABO3 perovskite oxides are conventional materials for electrochemical applications, in particular, SOFCs owing to their good reactivity, flexibility in terms of oxygen stoichiometry, and viability of A- and B-site cation substitutions. In most cases, the A-site cation is a combination of rare and alkaline earths for SOFC cathodes. However, the B-site cation is generally a transition metal, such as Mn, Fe, and Co. The atomic structure of perovskites consists in a 3D network of BO6 octahedra connected by apex, which are maintained by a cubic lattice of A atoms of coordination 12, as shown in Figure 2. The symmetries of ABO3 perovskites can theoretically be directly determined from the ionic radii of A and B cations, which govern strains inside the materials, since the size mismatch is responsible for the distinct stacking.
In the early stage, LaCoO3 (LCO113) [36,73] and LaMnO3 (LMO113) [36,74] based ABO3 perovskite oxides were utilized as a cathode material for SOFCs. These stoichiometric compositions, however, were found unsuitable in terms of electrochemical performance for a cathode. Usually, the increase of ionic conductivities is more influenced by Sr concentration at the A-site while the increase of the electronic conductivities is more influenced by Fe and Co concentration at the B-site. The substitution of Sr2+ for La3+ can enhance the oxygen ion transport kinetics due to an increase in the oxygen vacancy concentration in the perovskite structure [17]. Therefore, high electrochemical performance at the cathode could be achieved relative to undoped ABO3 perovskite oxides. In this sense, LSC113 [17,21,27,28,46,64,73,75,76] and La1−xSrxCo1−yFyO3−δ (LSCF113) [22,45,47,66,77,78,79,80,81,82,83,84,85,86] perovskite oxides are commonly used as these oxides are known to show oxygen deficiency and the oxygen nonstoichiometry (δ) affects the electrochemical properties, conductivity and lattice expansion of the materials.

3.1.1. Cobaltites

Epitaxial strain induced by lattice mismatch between a thin film and a substrate can effectively control the oxygen migration and oxygen defect formation in ABO3 perovskite oxides [64,87,88,89,90,91,92,93]. Using density functional theory (DFT) calculations, Han et al. [94] demonstrated that tensile strain increased the one-dimensional mobility of oxygen vacancy and adsorbed atomic oxygens migrations on the LCO113 surface, which can enhance the ORR kinetics. According to the authors, the maximum tensile strain for the enhanced ORR kinetics is 3%, which has a lower oxygen vacancy formation energy by 0.32 eV compared to the strain-free surface. Later, Mayeshiba and Morgan [90] predicted the oxygen migration barriers in LaMO3 (M = Sc, Ti, V, Cr, Mn, Fe, Co, Ni, and Ga) perovskite oxides under tensile strain via DFT calculations. They showed that tensile strain can significantly reduce the oxygen migration barriers, resulting in enhanced oxygen diffusion (one order of magnitude) at 500 °C compared with the diffusion under unstrained conditions. laO’ et al. [28] experimentally showed that tensile strained epitaxial La0.8Sr0.2CoO3−δ (LSC82) films exhibit significantly increased ORR activity compared to LSC82 bulk. According to the authors, the oxygen surface exchange coefficients (kq) of epitaxial LSC82 films are about two orders of magnitude higher than those of LSC82 bulk, which is attributed to the increased oxygen vacancy concentration in LSC82 thin film relative to LSC bulk. Using secondary ion mass spectroscopy (SIMS) measurements, Kubicek et al. [64] also investigated the effect of strain on the oxygen surface exchange kinetics of epitaxial LSC82 films. In this work, the epitaxial LSC82 films were synthesized on SrTiO3 (STO) and LaAlO3 (LAO) substrates to introduce tensile and compressive strain, respectively, into the LSC82 films, where the oxygen surface exchange kinetics was found to be faster in tensile strained films. The authors postulated that the lower formation energy of oxygen vacancy by tensile strain leads to the enhanced ORR kinetics. However, Crumlin et al. [27] did not observe the strong dependency of the oxygen surface exchange kinetics of epitaxial La0.6Sr0.4CoO3−δ (LSC64) films on strain. The authors attempted to control the degree of tensile strain in LSC64 films with various film thicknesses but the kq values of LSC64 films were thickness-independent and comparable with those of bulk LSC64 (kq can be approximated as k*). Similarly, epitaxial LCO thin films also showed little dependence of their surface exchange kinetics as a function of film thickness although the films exhibited higher ORR kinetics than LCO bulk [95]. In contrast, a clear trend of the thickness-dependent ORR activity in epitaxial LSC64 thin films was recently demonstrated by Lee et al. [96] (Figure 3a). According to this work, the oxygen surface exchange kinetics of LSC64 films can be enhanced as decreasing the film thickness until 50 nm, where strain relaxation occurs. The authors proposed that the increased ORR activity results from the increased oxygen vacancies in epitaxial LSC64 films due to the large tensile strain of high-quality films. Employing in situ impedance spectroscopy during thin film growth, Rupp et al. [97] also showed an enhancement in the oxygen surface exchange kinetics of tensile-strained epitaxial LSC64 films on (001)-oriented La0.95Sr0.05Ga0.95Mg0.05O3−δ substrates compared to unstrained films on yttria-stabilized zirconia (YSZ) substrates. Similar to the results by Lee et al. [96], the authors also found the same trend of enhanced oxygen surface exchange kinetics as film thickness decreases in epitaxial LSC64 films but they interpreted that this trend may be attributed to the capability of in-situ measurement as the films never exposed to any thermal cycling such as cooling from PLD preparation temperature or heating for impedance measurements. However, regardless of strain, epitaxial LSC113 thin films were found to have higher kq values compared to their bulk counterpart (Figure 3b).
It is known that lattice strain can tune the electronic structure and charge transfer in ABO3 oxides [98,99]. Using scanning tunneling spectroscopy and density functional theory calculations, Cai et al. [55] showed that the surface electronic state depends on the strain state of the epitaxial LSC82 thin films. In this work, STO and LAO substrates were utilized to induce tensile and compressive strains, respectively, into the LSC82 films, where tensile-strained LSC films exhibited a more increased electronic density of states (DOS) near the Fermi level compared to compressive strained LSC films. It was hypothesized that the increased DOS and the transition to the metal-like state are attributed to the easier formation of oxygen vacancies on tensile strained LSC films, which was supported by DFT calculations.
A recent study for a wide range of bulk perovskites demonstrated that the position of the oxygen 2p-band center with respect to the Fermi level is linearly correlated with the activation barrier of oxygen surface exchange, acting as a descriptor for the oxygen surface exchange kinetics [100]. According to this work, the formation energy of bulk oxygen vacancies decreases with increasing the oxygen 2p band center relative to the Fermi level, indicating the enhanced oxygen reduction reaction (Figure 4a). Later, Lee et al. [96] demonstrated that the oxygen 2p-band center relative to the Fermi level can be linearly correlated with strain state, where tensile strain results in the upshift of the oxygen 2p-band center (Figure 4b). According to the authors, As the oxygen 2p-band center also has a linear relationship with the vacancy formation energy in perovskite oxides [100], and thus the tensile strain is expected to lower the energy of vacancy formation due to an increase in the oxygen 2p-band center in the LSC113 films. Hong et al. [95] showed that in epitaxial LCO113 films, strain can be used to reduce the spin transition temperature and promote the occupation of higher spin states that weaken the Co-O bond, which results in an increase in the oxygen surface exchange kinetics. The authors also pointed out that the decreased Co-O bond strength due to the increased fractional occupancy of high-spin states can lead to the oxygen 2p-band center upshift relative to the Fermi energy. Recently, Zhu et al. [101] measured the ORR activities of polycrystalline LSC113 thin films with a wide range of Sr concentration (0 ≤ Sr ≤0.8). The authors found that with increasing the Sr content, the O 2p-band center can be close to the Fermi level, which favors the oxygen vacancy formation. However, they claimed that at Sr > 0.4, Co2+ becomes optimal for the surface instead of Co3+/Co4+ and may act as electron traps, which may decrease the electronic conductivity, leading to an overall poor ORR activity of LSC113 films. These results showed a different trend with the previously reported data [95,96] which showed enhanced ORR kinetics due to an upshift of the oxygen 2p-band center to the Fermi level.

3.1.2. Ferrites

Iron-based perovskites (LaFeO3, LFO113) are typically oxygen deficient materials [102,103]. Therefore, LFO113 has high oxygen mobility within the material [104] and is suitable for catalytic applications such as oxygen reduction and ion transport in SOFCs [105,106]. Substituting Sr for La on the A-site of LFO113 is widely used to promote oxygen defects, enhancing oxygen diffusion and surface exchange processes due to the charge neutrality [107]. However, the ORR activity of La1−xSrxFeO3 (LSF113) thin films is still insufficient for the use of oxygen electrocatalysis. For example, using electrochemical impedance spectroscopy (EIS) measurements, la O et al. [108] found that the oxygen surface exchange coefficients (kq) of La0.8Sr0.2FeO3 (LSF82) thin films are comparable to those of La0.8Sr0.2MnO3 (LSM82) thin films.
Compared to LSF113, LSCF113 exhibits better ionic conductivity, reasonable high-temperature stability, and catalytic activity [109,110,111,112,113]. However, there was no systematic study on the strain dependent ORR kinetics (Figure 5a). Baumann et al. [47] and Ingram et al. [85] reported the ORR kinetics of La0.6Sr0.4Co0.8Fe0.2O3−δ (LSCF6482) thin films (Figure 5b). The authors concluded that LSCF6428 films can enhance the ORR kinetics compared to bulk LSCF6428 due to the increased oxygen vacancies. So far, substantial efforts have been focused on tailoring the surface microstructure of LSCF113 thin films to enhance the ORR kinetics, based on the hypothesis that many grain boundaries on the surface of LSCF113 films may modify the oxygen bulk diffusion and catalytic properties to enhance the ORR kinetics. For instance, Prestat et al. [114] investigated the ORR kinetics of La0.52Sr0.48Co0.18Fe0.82O3−δ thin films with various thicknesses (Figure 5b). In this study, the polarization resistance of LSCF films was reduced by decreasing the film thickness, indicating that the number of grain boundaries on the surface increases. Consequently, the oxygen surface exchange and diffusion coefficients were significantly enhanced in LSCF thin films compared to bulk LSCF [84]. Angoua et al. [115] investigated the ORR kinetics of LSCF6428 thin films with a Ce0.8Gd0.2O1.9 (CGO) layer. Adding CGO to LSCF6428 films reduced the polarization resistance for LSCF-CGO films, which leads to a higher density of grain boundaries. Bagarinao et al. [116] discussed the oxygen surface exchange and diffusion kinetics of LSCF6428 thin films by comparing a nanograined microstructure film. The authors found that introducing grain boundaries on LSCF6428 films can enhance the oxygen surface exchange reactivity, which is in good agreement with previously reported data [115].
It was also shown that the ORR kinetics can be enhanced by using the LSCF113 thin film as an interfacial layer at the electrode-electrolyte interface. Dumaisnil et al. [120] investigated the ORR kinetics of adding LSCF6482 thin films between porous LSCF113 cathode and Gd-doped ceria (GDC) substrate. In this study, the addition of LSCF6482 thin films was found to result in lower polarization resistance, which may be attributed to the increased active area due to the enhanced adhesion between cathode/substrate. Jang et al. [121] also investigated the ORR kinetics by addition of porous nanoparticle network structured LSCF6428 thin film between the cathode and electrolyte. The cell with LSCF6428 films showed higher peak power density and smaller polarization resistance compared to the cell without LSCF6428 film. Later, the role of nanoparticle network structured LSCF6428 thin film, which is called nano-web structured LSCF6428 (NW-LSCF) was further discussed [122].

3.2. Ruddlesden-Popper (RP) Oxides

The general formula of Ruddlesden-Popper (RP) phases can be written as An+1BnO3n+1 [123]. The RP phases are comprised of n consecutive perovskite layers (ABO3) alternating with rock-salt layers (AO) along the crystallographic c-axis direction. Their formula can be represented by (AO)(ABO3)n, where n represents the number of connected layers of vertex sharing BO6 octahedra [124]. In the case of SOFCs, the n = 1 members of some RP series, notably those based on La2NiO4+δ (LNO214) are of particular interest. These RP oxides were shown to accommodate oxygen interstitials in the AO layers [10], leading to fast ion conduction and hence potential application as cathodes at intermediate temperatures. Recently, lanthanum strontium cobaltites have attracted much attention as a cathode material for SOFCs, which exhibited interestingly fast oxygen exchange kinetics when interfaced with its perovskite counterpart LSC113 [67,125]. In addition, (La,Sr)2CuO4 (LSCu214) has also demonstrated its availability as a promising cathode for SOFCs showing comparable oxygen surface exchange kinetics to that of LSCF113 thin films at intermediate temperatures [126]. Figure 6 presents the ideal tetragonal unit-cells for n = 1, which correspond to the stoichiometric compounds all of same space group, I4/mmm.
It is well-known that RP oxides exhibit strong anisotropic features as a result of the adjustment of two different structural units, i.e., ABO3 and AO within the lattice. While the B-site cations are coordinated by six oxygen anions, the B-O bond lengths are different due to the Jahn-Teller effect caused by the valence state of the B-site cations [127]. This results in two types of oxygen species in the BO6 octahedra, which are referred to as “apical” and “equatorial” oxygen. In the RP system, interstitial sites are located in the AO layer, where RP oxides can accommodate an excess oxygen as an interstitial oxygen defect. Oxygen vacancies can also be formed by appropriate doping in the system. Consequently, the oxygen transport properties of RP oxides can be strongly influenced by either oxygen interstitials or oxygen vacancies as discussed earlier.
Nickelates are interesting alternative materials to LSC113 for intermediate temperature SOFCs owing to its high oxygen surface exchange kinetics and oxygen transport properties [10,11,13,14,61,128,129,130,131,132,133]. However, due to difficulties in synthesizing epitaxial thin films of nickelates, few studies have examined the ORR kinetics of nickelates in a form of epitaxial films. Yamada et al. [134] first reported the oxygen electrocatalysis of epitaxial RP oxide thin films using (011)-oriented epitaxial Nd2NiO4+δ (NNO214) films on (001) YSZ substrates. According to this study, compression along the c -axis of NNO214 films reduces the oxygen surface exchange kinetics. Later, Tsvetkov et al. [135] compared the δ values of tensile- and compressive-strained NNO214 thin films, demonstrating that the larger δ in tensile-strained NNO214 can enhance the oxygen surface exchange kinetics relative to those of compressive-strained NNO214 thin films (Figure 7a). These results are in good agreement with the previous observation by Yamada et al. [134].
Among various RP oxides, LNO214 is the most widely studied member of the series. Kim et al. [136] proposed that epitaxial LNO214 thin films with a thickness of 300 nm may possess two different surface exchange rates as a result from the coexistence of two different film microstructures. Later, they demonstrated that the surface exchange reaction of polycrystalline LNO214 samples limits the electrode reaction, showing comparable kq values with those obtained previously for LNO214 bulk [132]. However, to evaluate the effect of strain on the ORR kinetics in RP oxides, high quality epitaxial thin films are required. Burriel and co-workers [61] synthesized (001)-oriented epitaxial LNO214 thin films on STO and NbGaO3 (NGO) substrates, which introduce tensile and compressive strain states, respectively, into the LNO214 films. They evaluated the oxygen diffusion and surface exchange kinetics for tensile- and compressive-strained LNO214 films, concluding that both compressive and tensile strain led to a reduction in the oxygen diffusivity in the first 175–200 nm of the film, whereas neither strain affected the oxygen surface exchange. However, Lee et al. [63] later demonstrated that tensile strain along the c-direction can lead to an increase in the driving force to form interstitial oxygen atoms in LNO214, resulting in enhanced oxygen surface exchange kinetics of (100)-oriented epitaxial LNO214 thin films (Figure 7b). In this study, the authors successfully controlled strain in LNO214 films by modulating the film thickness. The discrepancy in the effects of strain on oxygen surface exchange kinetics in LNO214 films may be attributed to the fact that the critical thickness required to maintain an acceptable strain state is fairly smaller than what Burriel et al. [61] used. However, the effect of strain on the ORR kinetics in RP oxides needs further investigation.
Similar to the correlation between the oxygen 2p-band center and ORR kinetics in ABO3 oxides, Lee et al. [126] also proposed that the oxygen 2p-band center relative to the Fermi energy can be an effective descriptor to predict the oxygen surface exchange as well as the activation energy for oxygen transport in RP oxides. In contrast to ABO3 oxides, the oxygen 2p-band center relative to the Fermi level was found to be inversely proportional to the oxygen surface exchange kinetics in RP oxides (Figure 8). According to the authors, the opposite slopes of the oxygen 2p-band center correlations between the RP oxides and the perovskite materials are due to the intrinsic mechanistic differences of their oxygen surface exchange kinetics and bulk anionic transport. The authors also found that the calculated interstitial formation energies of La2BO4+δ (B = Co, Ni, and Cu) strongly depend on the B cation in the sequence LCuO214 > LNO214 > LCO214, which corresponds to the trend of B-cation–dependent oxygen diffusivity. Recently, Xie et al. [137] also calculated the formation energy of oxygen interstitials in the same RP oxide systems, demonstrating that, regardless of δ, the formation energy of oxygen interstitials decreases with decreasing the atomic number of the B cation, which is in good agreement with the data reported by Lee et al. [126].
As discussed earlier, the anisotropy of oxygen migration is one of the intrinsic properties of RP oxides owing to its structural feature. Therefore, different oxygen diffusion and surface exchange kinetics along different crystallographic directions were reported as shown in Figure 9. Using NNO214 and Pr2NiO4+δ (PNO214) single crystalline oxides, Bassat et al. [128] observed a large anisotropy for the oxygen surface exchange kinetics for both oxides, where the k* values in the a-b plane are 1.5 orders of magnitude larger than those along the c-axis. Burriel et al. [61] reported that the oxygen diffusion and oxygen surface exchange kinetics in the a-b plane are faster than those along the c-direction in epitaxial LNO214 thin films. However, later the authors also evaluated the anisotropy for the oxygen surface exchange kinetics of LNO214 single crystalline, reporting the lack of anisotropy in the oxygen surface exchange values, which is attributed to the surface segregation process [133]. The anisotropy of the oxygen surface exchange kinetics was also observed in Sr-doped LNO214 films. Lee et al. [65] demonstrated that substituting Sr for lanthanum in LSNO214 thin films can result in the structural reorientation of the films because of the reduction in the surface energy of the (001) surface, reporting the anisotropic oxygen surface exchange kinetics.
La2-xSrxCuO4±δ (LSCu214) oxides are famous for their high Tc superconductivity [138,139,140]. Interestingly, the superconducting properties of LSCu214 depend on the Sr concentration and oxygen stoichiometry, which also strongly influence the kinetics of oxygen exchange and oxygen diffusion. Opila et al. reported the anisotropic oxygen transport properties and proposed the defect model for Sr-doped curpates [38,141]. Recently, using epitaxial LSCu214 thin films, the oxygen surface exchange kinetics were evaluated at intermediate temperatures [126]. In this study, the kq values of undoped La2CuO4 (LCuO214) thin films were found to be higher than those of Sr-doped La2CuO4 thin films (Figure 9a), which is attributed to that adding Sr content resulted in changes of the oxygen 2p-band center relative to the Fermi energy. More recently, Meyer and co-workers [142] revealed that the oxygen nonstoichiometry commonly reported for strained LSCu214 is mediated by the strain-modified surface exchange kinetics, rather than reduced thermodynamic oxygen formation energies. The authors also demonstrated that tensile-strained LSCu214 shows nearly an order of magnitude faster oxygen exchange rate in tensile-strained LSCu214 films compared to a compressively strained film.
La2-xSrxCoO4±δ (LSC214) oxides are widely used as a perovskite counterpart to promote the oxygen electrocatalysis of ABO3 perovskite such as LSC113 [33,68,125,143]. However, LSC214 showed very attractive transport and surface exchange properties at intermediate temperatures. For instance, La2CoO4+δ (LCO214) showed higher oxygen diffusivity and lower activation energy in the temperature range of 450–700 °C as compared to nickelates and cuprates [11,13,25]. Later, Lee et al. [126] calculated the interstitial formation energies of LCO214, LNO214, and LCuO214 and found the minimum energy for interstitial formation in LCO214. In addition, the authors also reported that the kq values of LSC214 thin films are comparable to LSC113 thin films, indicating a promising alternative cathode material to LSC113. Recently, Chen et al. [144] successfully fabricated LSC214 (Sr = 0.25) thin films grown in two different crystallographic orientations, i.e., (100) and (001), using (100) LaSrAlO4 (LSAO) and (001) SrTiO3 (STO) substrates. They demonstrated that the oxygen diffusivity along the a-b plane in the (100)-oriented LSC214 film was three orders of magnitude lower than that in the (001)-oriented LSC214 film (Figure 9b). Moreover, the D* value along the c-direction in the (001)-oriented LSC214 film was found to be one order of magnitude higher than that along the a-b plane in the (100)-oriented LSC214 film. However, no dependence of oxygen surface exchange kinetics was observed on the crystal orientation, which may be attributed to the strong Sr segregation at the surface overriding the effect of the structural anisotropy.

3.3. Oxide Heterostrcuture Interface

To achieve enhanced surface exchange kinetics and stability, recent efforts have been focused on developing advanced cathode materials based on MIECs with surface modification [27,33,66,68,76,82,86,125,143,145,146,147,148]. Oxide heterostructure interfaces, the combination of a Ruddlesden-Popper (RP, LSC214) layer on top of the perovskite LSC113, have shown remarkably high oxygen surface exchange kinetics [27,33,68,86,125,143,148,149]. Yashiro et al. reported ~1 order of magnitude enhancement in activity for the composite cathode screen-printed with the mixture of LSC64 and LSC214 [143]. Sase et al. [33] have also reported ~3 orders of magnitude higher oxygen surface exchange coefficient (k*) at the interfacial region between polycrystalline LSC64 and LSC214 compared to their bulk value [21]. Later, Crumlin et al. [27,148] utilized epitaxial LSC113 and LSC214 thin films, achieving higher oxygen surface exchange coefficients of LSC113. Recently, Zhao et al. [150] also observed ~10 times higher kq values in heterostructure LSC113/LSC214 thin films than those in LSC113, which is attributed to increased number of active oxygen vacancies that manipulate the oxygen 2p-band center.
A recent study using Coherent Brag Rod Analysis (COBRA) revealed the atomic structure and concentrations of the (001)-oriented LSC113 thin film on a STO substrate, which shows strontium (Sr) segregation toward the LSC113 surface and Sr depletion near the interface between LSC113 and STO [151]. It was also reported that heating the (001)-oriented LSC113 surface leads to the formation of surface Sr-enriched particles upon annealing while the LSC214-decorated LSC113 surface chemistry is stable upon heating [152]. More recently, Lee et al. [86] compared the time-dependent surface exchange kinetics and stability of epitaxial LSCF113 and LSC113 thin films with LSC214 surface decoration. Using density functional theory (DFT) computation, the authors revealed greater surface Sr segregation for LSCF113, which is predicted to have an SrO termination, than LSC113, which is predicted to have a less Sr enriched (La0.25Sr0.75)O termination. Furthermore, the authors pointed out a lower energy gain to move Sr from LSCF113 into LSC214 relative to the LSC214-LSC113 surface (Figure 10). These observations have suggested that the surface decoration can modulate the surface Sr segregation and the surface phase stability, which can greatly influence the oxygen surface exchange kinetics and the surface stability in LSC113 and LSCF113. Therefore, understanding the surface decoration effect on the surface chemistry of perovskites is critical and other potential surface modification materials need to be investigated to design highly active and stable cathodes for SOFCs.
Similar to the LSC113 surface modification with LSC214, several studies have also reported the enhanced surface electrocatalytic activity of porous LSCF cathodes with surface decoration [82,147,153,154]. For example, depositing thin La0.85Sr0.15MnO3−δ coatings on porous LSCF6428 electrodes using an infiltration process, Lynch et al. [82] showed the enhanced surface electrocatalytic activity of decorated LSCF113 cathodes upon polarization. Vast majority of research has been performed on porous LSCF113 electrodes, which lead to ambiguous structure and geometry, and therefore the physical origin responsible for enhanced cathodic performance associated with surface decoration of perovskites is not yet completely understood. However, Lee et al. [66] recently demonstrated using new forms of heterostructure oxide interfaces—(1) LSCF6428 thin films with a single layer decoration of mixed LSC214 and LSC113 and (2) LSCF6428 thin films with a double layer decoration of stacked LSC214 and LSC113 – significantly enhanced ORR activity of LSCF6428. More recently, Chen et al. [155] showed the improved electrocatalytic activity and durability of LSCF6428 thin films by coating PrO2. The authors claimed that PrO2 coating facilitates the rate of oxygen dissociation due to increased oxygen vacancies. Figure 11 shows the influence of heterointerfaces on k* for LSC113 and LSCF113.

4. Conclusions

Developing intermediate temperature SOFCs requires discovering new cathode materials with faster ORR kinetics. MIEC oxides such as ABO3 oxides and RP oxides are promising alternative cathode materials to LSM113, pure electronic conductor. Recently, introducing oxide thin films and oxide heterostructure interfaces has been recognized as an extremely effective means to control the ORR kinetics of MIEC oxides, opening a plethora of new possibilities useful for intermediate temperature SOFCs. We have briefly overviewed how epitaxial strain can influence the oxygen surface exchange kinetics in ABO3 and RP oxide thin films. It has been pointed out that strain can tune the oxygen 2p-band center relative to the Fermi energy as a result of the modulated concentration of oxygen defects. In the case of RP oxides, epitaxial thin films offer a direct way to evaluate the intrinsic anisotropy of oxygen diffusion and surface exchange kinetics. Lastly, oxide heterointerfaces can be used to improve the ORR kinetics and stability of ABO3 oxides. This review has given a comprehensive overview of the current state of research. However, a full understanding of the effect of strain and crystallographic orientation on the ORR kinetics in ABO3 and RP oxides is still being developed for intermediate SOFCs. In addition, understanding the surface decoration effect on the surface chemistry of perovskites and other potential surface modification materials still need to be investigated to design highly active and stable cathodes for SOFCs.

Author Contributions

D.L. conceived and coordinated the whole idea of writing this paper. G.Y. and W.J. identified the most active and promising research topics in the field under review and analyzed the selected scientific literature. S.-J.A. consulted the relevant studies and references. All the authors contributed to the writing of the manuscript.

Funding

This research was funded by Advanced Support Program for Innovative Research Excellence-I funding (#15540-18-47811) provided by the Office of the Vice President for Research at the University of South Carolina.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Adler, S.B. Factors governing oxygen reduction in solid oxide fuel cell cathodes. Chem. Rev. 2004, 104, 4791–4843. [Google Scholar] [CrossRef] [PubMed]
  2. Jacobson, A.J. Materials for solid oxide fuel cells. Chem. Mater. 2010, 22, 660–674. [Google Scholar] [CrossRef]
  3. Shao, Z.P.; Haile, S.M. A high-performance cathode for the next generation of solid-oxide fuel cells. Nature 2004, 431, 170–173. [Google Scholar] [CrossRef] [PubMed]
  4. Steele, B.C.H.; Heinzel, A. Materials for fuel-cell technologies. Nature 2001, 414, 345–352. [Google Scholar] [CrossRef]
  5. Wachsman, E.D.; Lee, K.T. Lowering the temperature of solid oxide fuel cells. Science 2011, 334, 935–939. [Google Scholar] [CrossRef] [PubMed]
  6. Kawada, T.; Suzuki, J.; Sase, M.; Kaimai, A.; Yashiro, K.; Nigara, Y.; Mizusaki, J.; Kawamura, K.; Yugami, H. Determination of oxygen vacancy concentration in a thin film of La0.6Sr0.4CoO3−δ by an electrochemical method. J. Electrochem. Soc. 2002, 149, E252–E259. [Google Scholar] [CrossRef]
  7. Mizusaki, J.; Mima, Y.; Yamauchi, S.; Fueki, K.; Tagawa, H. Nonstoichiometry of the perovskite-type oxides La1−xSrxCoO3−δ. J. Solid State Chem. 1989, 80, 102–111. [Google Scholar] [CrossRef]
  8. De Souza, R.A.; Kilner, J.A. Oxygen transport in La1−xSrxMn1−yCoyO3±δ perovskites: Part II. Oxygen surface exchange. Solid State Ion. 1999, 126, 153–161. [Google Scholar] [CrossRef]
  9. De Souza, R.A.; Kilner, J.A. Oxygen transport in La1−xSrxMn1−yCoyO3±δ perovskites: Part I. Oxygen tracer diffusion. Solid State Ion. 1998, 106, 175–187. [Google Scholar] [CrossRef]
  10. Bassat, J.M.; Odier, P.; Villesuzanne, A.; Marin, C.; Pouchard, M. Anisotropic ionic transport properties in La2NiO4+δ single crystals. Solid State Ion. 2004, 167, 341–347. [Google Scholar] [CrossRef]
  11. Boehm, E.; Bassat, J.M.; Steil, M.C.; Dordor, P.; Mauvy, F.; Grenier, J.C. Oxygen transport properties of La2Ni1−xCuxO4+δ mixed conducting oxides. Solid State Sci. 2003, 5, 973–981. [Google Scholar] [CrossRef]
  12. Lee, D.; Lee, H. Controlling oxygen mobility in Ruddlesden–Popper oxides. Materials 2017, 10, 368. [Google Scholar] [CrossRef]
  13. Boehm, E.; Bassat, J.M.; Dordor, P.; Mauvy, F.; Grenier, J.C.; Stevens, P. Oxygen diffusion and transport properties in non-stoichiometric Ln2−xNiO4+δ oxides. Solid State Ion. 2005, 176, 2717–2725. [Google Scholar] [CrossRef]
  14. Kilner, J.A.; Shaw, C.K.M. Mass transport in La2Ni1−xCoxO4+δ oxides with the K2NiF4 structure. Solid State Ion. 2002, 154–155, 523–527. [Google Scholar] [CrossRef]
  15. Ishigaki, T.; Yamauchi, S.; Kishio, K.; Mizusaki, J.; Fueki, K. Diffusion of oxide ion vacancies in perovskite-type oxides. J. Solid State Chem. 1988, 73, 179–187. [Google Scholar] [CrossRef]
  16. Ishigaki, T.; Yamauchi, S.; Mizusaki, J.; Fueki, K.; Tamura, H. Tracer diffusion-coefficient of oxide ions in LaCoO3 single-crystal. J. Solid State Chem. 1984, 54, 100–107. [Google Scholar] [CrossRef]
  17. Mizusaki, J. Nonstoichiometry, diffusion, and electrical-properties of perovskite-type oxide electrode materials. Solid State Ion. 1992, 52, 79–91. [Google Scholar] [CrossRef]
  18. Routbort, J.L.; Doshi, R.; Krumpelt, M. Oxygen tracer diffusion in La1−xSrxCoO3. Solid State Ion. 1996, 90, 21–27. [Google Scholar] [CrossRef]
  19. VanDoorn, R.H.E.; Fullarton, I.C.; De Souza, R.A.; Kilner, J.A.; Bouwmeester, H.J.M.; Burggraaf, A.J. Surface oxygen exchange of La0.3Sr0.7CoO3−δ. Solid State Ion. 1997, 96, 1–7. [Google Scholar] [CrossRef]
  20. Yildiz, B. “Stretching” the energy landscape of oxides-Effects on electrocatalysis and diffusion. MRS Bull. 2014, 39, 147–156. [Google Scholar] [CrossRef]
  21. Berenov, A.V.; Atkinson, A.; Kilner, J.A.; Bucher, E.; Sitte, W. Oxygen tracer diffusion and surface exchange kinetics in La0.6Sr0.4CoO3−δ. Solid State Ion. 2010, 181, 819–826. [Google Scholar] [CrossRef]
  22. Carter, S.; Selcuk, A.; Chater, R.J.; Kajda, J.; Kilner, J.A.; Steele, B.C.H. Oxygen transport in selected nonstoichiometric perovskite-structure oxides. Solid State Ion. 1992, 53, 597–605. [Google Scholar] [CrossRef]
  23. Lane, J.A.; Benson, S.J.; Waller, D.; Kilner, J.A. Oxygen transport in La0.6Sr0.4Co0.2Fe0.8O3−δ. Solid State Ion. 1999, 121, 201–208. [Google Scholar] [CrossRef]
  24. Wang, L.; Merkle, R.; Maier, J.; Acartürk, T.; Starke, U. Oxygen tracer diffusion in dense Ba0.5Sr0.5Co0.8Fe0.2O3−δ films. Appl. Phys. Lett. 2009, 94, 071908. [Google Scholar] [CrossRef]
  25. Munnings, C.N.; Skinner, S.J.; Amow, G.; Whitfield, P.S.; Davidson, I.J. Oxygen transport in the La2Ni1−xCoxO4+δ system. Solid State Ion. 2005, 176, 1895–1901. [Google Scholar] [CrossRef]
  26. Baumann, F.S.; Maier, J.; Fleig, J. The polarization resistance of mixed conducting SOFC cathodes: A comparative study using thin film model electrodes. Solid State Ion. 2008, 179, 1198–1204. [Google Scholar] [CrossRef]
  27. Crumlin, E.J.; Ahn, S.J.; Lee, D.; Mutoro, E.; Biegalski, M.D.; Christen, H.M.; Shao-Horn, Y. Oxygen electrocatalysis on epitaxial La0.6Sr0.4CoO3−δ perovskite thin films for solid oxide fuel cells. J. Electrochem. Soc. 2012, 159, F219–F225. [Google Scholar] [CrossRef]
  28. LaO’, G.J.; Ahn, S.J.; Crumlin, E.; Orikasa, Y.; Biegalski, M.D.; Christen, H.M.; Shao-Horn, Y. Catalytic activity enhancement for oxygen reduction on epitaxial perovskite thin films for solid-oxide fuel cells. Angew. Chem. 2010, 49, 5344–5347. [Google Scholar]
  29. Santiso, J.; Burriel, M. Deposition and characterisation of epitaxial oxide thin films for SOFCs. J. Solid State Electrochem. 2011, 15, 985–1006. [Google Scholar] [CrossRef]
  30. Lutchyn, R.M.; Sau, J.D.; Das Sarma, S. Majorana fermions and a topological phase transition in semiconductor-superconductor heterostructures. Phys. Rev. Lett. 2010, 105, 077001. [Google Scholar] [CrossRef]
  31. Smadici, Ş.; Abbamonte, P.; Bhattacharya, A.; Zhai, X.; Jiang, B.; Rusydi, A.; Eckstein, J.N.; Bader, S.D.; Zuo, J.M. Electronic reconstruction at SrMnO3-LaMnO3 superlattice interfaces. Phys. Rev. Lett. 2007, 99, 196404. [Google Scholar] [CrossRef]
  32. Chakhalian, J.; Freeland, J.W.; Srajer, G.; Strempfer, J.; Khaliullin, G.; Cezar, J.C.; Charlton, T.; Dalgliesh, R.; Bernhard, C.; Cristiani, G.; et al. Magnetism at the interface between ferromagnetic and superconducting oxides. Nat. Phys. 2006, 2, 244–248. [Google Scholar] [CrossRef]
  33. Sase, M.; Hermes, F.; Yashiro, K.; Sato, K.; Mizusaki, J.; Kawada, T.; Sakai, N.; Yokokawa, H. Enhancement of oxygen surface exchange at the hetero-interface of (La,Sr)CoO3/(La,Sr)2CoO4 with PLD-layered films. J. Electrochem. Soc. 2008, 155, B793–B797. [Google Scholar] [CrossRef]
  34. Chroneos, A.; Vovk, R.V.; Goulatis, I.L.; Goulatis, L.I. Oxygen transport in perovskite and related oxides: A brief review. J. Alloys Compd. 2010, 494, 190–195. [Google Scholar] [CrossRef]
  35. Kharton, V.V.; Viskup, A.P.; Kovalevsky, A.V.; Naumovich, E.N.; Marques, F.M.B. Ionic transport in oxygen-hyperstoichiometric phases with K2NiF4-type structure. Solid State Ion. 2001, 143, 337–353. [Google Scholar] [CrossRef]
  36. Minh, N.Q. Ceramic fuel-cells. J. Am. Ceram. Soc. 1993, 76, 563–588. [Google Scholar] [CrossRef]
  37. Minervini, L.; Grimes, R.W.; Kilner, J.A.; Sickafus, K.E. Oxygen migration in La2NiO4+δ. J. Mater. Chem. 2000, 10, 2349–2354. [Google Scholar] [CrossRef]
  38. Opila, E.J.; Tuller, H.L.; Wuensch, B.J.; Maier, J. Oxygen tracer diffusion in La2−xSrxCuO4−δ single-crystals. J. Am. Ceram. Soc. 1993, 76, 2363–2369. [Google Scholar] [CrossRef]
  39. Horita, T.; Yamaji, K.; Sakai, N.; Xiong, Y.; Yokokawa, H.; Kawada, T.J.I. Application of novel SIMS technique for imaging the active sites of oxygen reduction at the SOFC cathode/electrolyte interfaces. Ionics 2002, 8, 108–117. [Google Scholar] [CrossRef]
  40. Horita, T.; Yamaji, K.; Sakai, N.; Yokokawa, H.; Weber, A.; Ivers-Tiffee, E. Oxygen reduction mechanism at porous La1−xSrxCoO3−δ cathodes/La0.8Sr0.2Ga0.8Mg0.2O2.8 electrolyte interface for solid oxide fuel cells. Electrochim. Acta 2001, 46, 1837–1845. [Google Scholar] [CrossRef]
  41. Kamata, H.; Hosaka, A.; Mizusaki, J.; Tagawa, H. High temperature electrocatalytic properties of the SOFC air electrode La0.8Sr0.2MnO3/YSZ. Solid State Ion. 1998, 106, 237–245. [Google Scholar] [CrossRef]
  42. Opitz, A.K.; Schintlmeister, A.; Hutter, H.; Fleig, J. Visualization of oxygen reduction sites at Pt electrodes on YSZ by means of 18O tracer incorporation: The width of the electrochemically active zone. Phys. Chem. Chem. Phys. 2010, 12, 12734–12745. [Google Scholar] [CrossRef]
  43. Østergård, M.J.L.; Mogensen, M. ac Impedance study of the oxygen reduction mechanism on La1−xSrxMnO3 in solid oxide fuel cells. Electrochim. Acta 1993, 38, 2015–2020. [Google Scholar] [CrossRef]
  44. Robertson, N.L.; Michaels, J.N. Oxygen exchange on platinum electrodes in zirconia cells: Location of electrochemical reaction sites. J. Electrochem. Soc. 1990, 137, 129–135. [Google Scholar] [CrossRef]
  45. Bouwmeester, H.J.M.; Kruidhof, H.; Burggraaf, A.J. Importance of the surface exchange kinetics as rate-limiting step in oxygen permeation through mixed-conducting oxides. Solid State Ion. 1994, 72, 185–194. [Google Scholar] [CrossRef]
  46. Van der Haar, L.M.; den Otter, M.W.; Morskate, M.; Bouwmeester, H.J.M.; Verweij, H. Chemical diffusion and oxygen surface transfer of La1−xSrxCoO3−δ studied with electrical conductivity relaxation. J. Electrochem. Soc. 2002, 149, J41–J46. [Google Scholar] [CrossRef]
  47. Baumann, F.S.; Fleig, J.; Habermeier, H.U.; Maier, J. Impedance spectroscopic study on well-defined (La,Sr)(Co,Fe)O3−δ model electrodes. Solid State Ion. 2006, 177, 1071–1081. [Google Scholar] [CrossRef]
  48. Jamnik, J.; Maier, J. Generalised equivalent circuits for mass and charge transport: Chemical capacitance and its implications. Phys. Chem. Chem. Phys. 2001, 3, 1668–1678. [Google Scholar] [CrossRef]
  49. Yang, Y.L.; Chen, C.L.; Chen, S.Y.; Chu, C.W.; Jacobson, A.J. Impedance studies of oxygen exchange on dense thin film electrodes of La0.5Sr0.5CoO3−δ. J. Electrochem. Soc. 2000, 147, 4001–4007. [Google Scholar] [CrossRef]
  50. Endo, A.; Fukunaga, H.; Wen, C.; Yamada, K. Cathodic reaction mechanism of dense La0.6Sr0.4CoO3 and La0.81Sr0.09MnO3 electrodes for solid oxide fuel cells. Solid State Ion. 2000, 135, 353–358. [Google Scholar] [CrossRef]
  51. Yang, Y.L.; Jacobson, A.J.; Chen, C.L.; Luo, G.P.; Ross, K.D.; Chu, C.W. Oxygen exchange kinetics on a highly oriented La0.5Sr0.5CoO3−δ thin film prepared by pulsed-laser deposition. Appl. Phys. Lett. 2001, 79, 776–778. [Google Scholar] [CrossRef]
  52. Sase, M.; Ueno, D.; Yashiro, K.; Kaimai, A.; Kawada, T.; Mizusaki, J. Interfacial reaction and electrochemical properties of dense (La,Sr)CoO3−δ cathode on YSZ (100). J. Phys. Chem. Solids 2005, 66, 343–348. [Google Scholar] [CrossRef]
  53. Maier, J. Physical chemistry of ionic materials: Ions and electrons in solids. Wiley 2004. [Google Scholar]
  54. Fleig, J.; Maier, J. The polarization of mixed conducting SOFC cathodes: Effects of surface reaction coefficient, ionic conductivity and geometry. J. Eur. Ceram. Soc. 2004, 24, 1343–1347. [Google Scholar] [CrossRef]
  55. Cai, Z.; Kuru, Y.; Han, J.W.; Chen, Y.; Yildiz, B. Surface electronic structure transitions at high temperature on perovskite oxides: The case of strained La0.8Sr0.2CoO3 thin films. J. Am. Chem. Soc. 2011, 133, 17696–17704. [Google Scholar] [CrossRef]
  56. Donner, W.; Chen, C.; Liu, M.; Jacobson, A.J.; Lee, Y.L.; Gadre, M.; Morgan, D. Epitaxial strain-induced chemical ordering in La0.5Sr0.5CoO3−δ films on SrTiO3. Chem. Mater. 2011, 23, 984–988. [Google Scholar] [CrossRef]
  57. Garcia-Barriocanal, J.; Rivera-Calzada, A.; Varela, M.; Sefrioui, Z.; Iborra, E.; Leon, C.; Pennycook, S.J.; Santamaria, J. Colossal ionic conductivity at interfaces of epitaxial ZrO2:Y2O3/SrTiO3 heterostructures. Science 2008, 321, 676–680. [Google Scholar] [CrossRef]
  58. Schichtel, N.; Korte, C.; Hesse, D.; Janek, J. Elastic strain at interfaces and its influence on ionic conductivity in nanoscaled solid electrolyte thin films—Theoretical considerations and experimental studies. Phys. Chem. Chem. Phys. 2009, 11, 3043–3048. [Google Scholar] [CrossRef]
  59. Wang, Z.W.; Shu, D.J.; Wang, M.; Ming, N.B. Diffusion of oxygen vacancies on a strained rutile TiO2 (110) surface. Phys. Rev. B 2010, 82, 165309. [Google Scholar] [CrossRef]
  60. Kushima, A.; Yip, S.; Yildiz, B. Competing strain effects in reactivity of LaCoO3 with oxygen. Phys. Rev. B 2010, 82, 115435. [Google Scholar] [CrossRef]
  61. Burriel, M.; Garcia, G.; Santiso, J.; Kilner, J.A.; Chater, R.J.; Skinner, S.J. Anisotropic oxygen diffusion properties in epitaxial thin films of La2NiO4+δ. J. Mater. Chem. 2008, 18, 416–422. [Google Scholar] [CrossRef]
  62. Ji, H.I.; Hwang, J.; Yoon, K.J.; Son, J.W.; Kim, B.K.; Lee, H.W.; Lee, J.H. Enhanced oxygen diffusion in epitaxial lanthanum–strontium–cobaltite thin film cathodes for micro solid oxide fuel cells. Energy Environ. Sci. 2013, 6, 116–120. [Google Scholar] [CrossRef]
  63. Lee, D.; Grimaud, A.; Crumlin, E.J.; Mezghani, K.; Habib, M.A.; Feng, Z.; Hong, W.T.; Biegalski, M.D.; Christen, H.M.; Shao-Horn, Y. Strain influence on the oxygen electrocatalysis of the (100)-oriented epitaxial La2NiO4+δ thin films at elevated temperatures. J. Phys. Chem. C 2013, 117, 18789–18795. [Google Scholar] [CrossRef]
  64. Kubicek, M.; Cai, Z.; Ma, W.; Yildiz, B.; Hutter, H.; Fleig, J. Tensile lattice strain accelerates oxygen surface exchange and diffusion in La1–xSrxCoO3−δ thin films. ACS Nano 2013, 7, 3276–3286. [Google Scholar] [CrossRef]
  65. Lee, D.; Lee, Y.L.; Grimaud, A.; Hong, W.T.; Biegalski, M.D.; Morgan, D.; Shao-Horn, Y. Strontium influence on the oxygen electrocatalysis of La2−xSrxNiO4±δ (0.0 ≤ xSr ≤ 1.0) thin films. J. Mater. Chem. A 2014, 2, 6480–6487. [Google Scholar] [CrossRef]
  66. Lee, D.; Lee, Y.L.; Wang, X.R.; Morgan, D.; Shao-Horn, Y. Enhancement of oxygen surface exchange on epitaxial La0.6Sr0.4Co0.2Fe0.8O3−δ thin films using advanced heterostructured oxide interface engineering. MRS Commun. 2016, 6, 204–209. [Google Scholar] [CrossRef]
  67. Sase, M.; Hermes, F.; Nakamura, T.; Yashiro, K.; Sato, K.; Mizusaki, J.; Kawada, T.; Sakai, N.; Yamaji, K.; Horita, T.; et al. Promotion of oxygen surface reaction at the hetero-interface of (La,Sr)CoO3/(La,Sr)2CoO4. ECS Trans. 2007, 7, 1055–1060. [Google Scholar]
  68. Yashiro, K.; Nakamura, T.; Sase, M.; Hermes, F.; Sato, K.; Kawada, T.; Mizusaki, J. Electrode performance at hetero-interface of perovskite-related oxides, (La,Sr)CoO3−δ/(La,Sr)2CoO4−δ. ECS Trans. 2007, 7, 1287–1292. [Google Scholar]
  69. Schlom, D.G.; Haeni, J.H.; Lettieri, J.; Theis, C.D.; Tian, W.; Jiang, J.C.; Pan, X.Q. Oxide nano-engineering using MBE. Mater. Sci. Eng. B 2001, 87, 282–291. [Google Scholar] [CrossRef]
  70. Elam, J.W.; Groner, M.D.; George, S.M. Viscous flow reactor with quartz crystal microbalance for thin film growth by atomic layer deposition. Rev. Sci. Instrum. 2002, 73, 2981–2987. [Google Scholar] [CrossRef]
  71. Lowndes, D.H.; Geohegan, D.B.; Puretzky, A.A.; Norton, D.P.; Rouleau, C.M. Synthesis of novel thin-film materials by pulsed laser deposition. Science 1996, 273, 898–903. [Google Scholar] [CrossRef]
  72. Alfonso, E.; Olaya, J.; Cubillos, G. Thin film growth through sputtering technique and its applications. In Crystallization—Science and Technology; IntechOpen: London, UK, 2012. [Google Scholar]
  73. Adler, S.B. Mechanism and kinetics of oxygen reduction on porous La1−xSrxCoO3−δ electrodes. Solid State Ion. 1998, 111, 125–134. [Google Scholar] [CrossRef]
  74. Jiang, Y.; Wang, S.Z.; Zhang, Y.H.; Yan, J.W.; Li, W.Z. Electrochemical reduction of oxygen on a strontium doped lanthanum manganite electrode. Solid State Ion. 1998, 110, 111–119. [Google Scholar] [CrossRef]
  75. Van Doorn, R.H.E.; Burggraaf, A.J. Structural aspects of the ionic conductivity of La1−xSrxCoO3−δ. Solid State Ion. 2000, 128, 65–78. [Google Scholar] [CrossRef]
  76. Lee, D.; Lee, Y.L.; Grimaud, A.; Hong, W.T.; Biegalski, M.D.; Morgan, D.; Shao-Horn, Y. Enhanced oxygen surface exchange kinetics and stability on epitaxial La0.8Sr0.2CoO3−δ thin films by La0.8Sr0.2MnO3−δ decoration. J. Phys. Chem. C 2014, 118, 14326–14334. [Google Scholar] [CrossRef]
  77. Esquirol, A.; Brandon, N.P.; Kilner, J.A.; Mogensen, M. Electrochemical characterization of La0.6Sr0.4Co0.2Fe0.8O3 cathodes for intermediate-temperature SOFCs. J. Electrochem. Soc. 2004, 151, A1847–A1855. [Google Scholar] [CrossRef]
  78. Geary, T.C.; Lee, D.; Shao-Horn, Y.; Adler, S.B. Nonlinear impedance analysis of La0.4Sr0.6Co0.2Fe0.8O3−δ thin film oxygen electrodes. J. Electrochem. Soc. 2016, 163, F1107–F1114. [Google Scholar] [CrossRef]
  79. Jiang, S.P. A comparison of O2 reduction reactions on porous (La,Sr)MnO3 and (La,Sr)(Co,Fe)O3 electrodes. Solid State Ion. 2002, 146, 1–22. [Google Scholar] [CrossRef]
  80. Katsuki, M.; Wang, S.; Dokiya, M.; Hashimoto, T. High temperature properties of La0.6Sr0.4Co0.8Fe0.2O3−δ oxygen nonstoichiometry and chemical diffusion constant. Solid State Ion. 2003, 156, 453–461. [Google Scholar] [CrossRef]
  81. Kuhn, M.; Fukuda, Y.; Hashimoto, S.; Sato, K.; Yashiro, K.; Mizusaki, J. Oxygen nonstoichiometry and thermo-chemical stability of perovskite-type La0.6Sr0.4Co1−yFeyO3−δ (y = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) materials. J. Electrochem. Soc. 2013, 160, F34–F42. [Google Scholar] [CrossRef]
  82. Lynch, M.E.; Yang, L.; Qin, W.T.; Choi, J.J.; Liu, M.F.; Blinn, K.; Liu, M.L. Enhancement of La0.6Sr0.4Co0.2Fe0.8O3−δ durability and surface electrocatalytic activity by La0.85Sr0.15MnO3±δ investigated using a new test electrode platform. Energy Environ. Sci. 2011, 4, 2249–2258. [Google Scholar] [CrossRef]
  83. Tai, L.W.; Nasrallah, M.M.; Anderson, H.U.; Sparlin, D.M.; Sehlin, S.R. Structure and electrical-properties of La1−xSrxCo1−yFeyO3 Part 1. The system La0.8Sr0.2Co1−yFeyO3. Solid State Ion. 1995, 76, 259–271. [Google Scholar] [CrossRef]
  84. Steele, B.C.H.; Bae, J.M. Properties of La0.6Sr0.4Co0.2Fe0.8O3x (LSCF) double layer cathodes on gadolinium-doped cerium oxide (CGO) electrolytes: II. Role of oxygen exchange and diffusion. Solid State Ion. 1998, 106, 255–261. [Google Scholar] [CrossRef]
  85. Ingram, B.J.; Eastman, J.A.; Chang, K.C.; Kim, S.K.; Fister, T.T.; Perret, E.; You, H.; Baldo, P.M.; Fuoss, P.H. In situ X-ray studies of oxygen surface exchange behavior in thin film La0.6Sr0.4Co0.2Fe0.8O3−δ. Appl. Phys. Lett. 2012, 101, 051603. [Google Scholar] [CrossRef]
  86. Lee, D.; Lee, Y.L.; Hong, W.T.; Biegalski, M.D.; Morgan, D.; Shao-Horn, Y. Oxygen surface exchange kinetics and stability of (La,Sr)2CoO4±δ/La1−xSrxMO3−δ (M = Co and Fe) hetero-interfaces at intermediate temperatures. J. Mater. Chem. A 2015, 3, 2144–2157. [Google Scholar] [CrossRef]
  87. Cho, S.Y.; Chung, Y.C.; Ahn, K.; Lee, J.H.; Kim, B.K.; Kim, H. Oxygen transport in epitaxial La0.875Sr0.125CoO3−δ thin-film cathodes for solid oxide fuel cells: Roles of anisotropic strain. Scr. Mater. 2016, 115, 141–144. [Google Scholar] [CrossRef]
  88. Gan, L.Y.; Akande, S.O.; Schwingenschlogl, U. Anisotropic O vacancy formation and diffusion in LaMnO3. J. Mater. Chem. A 2014, 2, 19733–19737. [Google Scholar] [CrossRef]
  89. Kushima, A.; Yildiz, B. Role of lattice strain and defect chemistry on the oxygen vacancy migration at the (8.3% Y2O3-ZrO2) / SrTiO3 hetero-interface: A first principles study. ECS Trans. 2009, 25, 1599–1609. [Google Scholar]
  90. Mayeshiba, T.; Morgan, D. Strain effects on oxygen migration in perovskites. Phys. Chem. Chem. Phys. 2015, 17, 2715–2721. [Google Scholar] [CrossRef]
  91. Petrie, J.R.; Jeen, H.; Barron, S.C.; Meyer, T.L.; Lee, H.N. Enhancing perovskite electrocatalysis through strain tuning of the oxygen deficiency. J. Am. Chem. Soc. 2016, 138, 7252–7255. [Google Scholar] [CrossRef]
  92. Petrie, J.R.; Mitra, C.; Jeen, H.; Choi, W.S.; Meyer, T.L.; Reboredo, F.A.; Freeland, J.W.; Eres, G.; Lee, H.N. Strain control of oxygen vacancies in epitaxial strontium cobaltite films. Adv. Funct. Mater. 2016, 26, 1564–1570. [Google Scholar] [CrossRef]
  93. Xu, T.; Shimada, T.; Araki, Y.; Wang, J.; Kitamura, T. Defect-strain engineering for multiferroic and magnetoelectric properties in epitaxial (110) ferroelectric lead titanate. Phys. Rev. B 2015, 92, 104106. [Google Scholar] [CrossRef]
  94. Han, J.W.; Yildiz, B. Enhanced one dimensional mobility of oxygen on strained LaCoO3(001) surface. J. Mater. Chem. 2011, 21, 18983–18990. [Google Scholar] [CrossRef]
  95. Hong, W.T.; Gadre, M.; Lee, Y.L.; Biegalski, M.D.; Christen, H.M.; Morgan, D.; Shao-Horn, Y. Tuning the spin state in LaCoO3 thin films for enhanced high-temperature oxygen electrocatalysis. J. Phys. Chem. Lett. 2013, 4, 2493–2499. [Google Scholar] [CrossRef]
  96. Lee, D.; Jacobs, R.; Jee, Y.; Seo, A.; Sohn, C.; Ievlev, A.V.; Ovchinnikova, O.S.; Huang, K.; Morgan, D.; Lee, H.N. Stretching epitaxial La0.6Sr0.4CoO3−δ for fast oxygen reduction. J. Phys. Chem. C 2017, 121, 25651–25658. [Google Scholar] [CrossRef]
  97. Rupp, G.M.; Kubicek, M.; Opitz, A.K.; Fleig, J. In situ impedance analysis of oxygen exchange on growing La0.6Sr0.4CoO3−δ thin films. ACS Appl. Energy Mater. 2018, 1, 4522–4535. [Google Scholar] [CrossRef]
  98. Hou, F.; Cai, T.Y.; Ju, S.; Shen, M.R. Half-metallic ferromagnetism via the interface electronic reconstruction in LaAlO3/SrMnO3 nanosheet superlattices. ACS Nano 2012, 6, 8552–8562. [Google Scholar] [CrossRef]
  99. Rata, A.D.; Herklotz, A.; Nenkov, K.; Schultz, L.; Dörr, K. Strain-induced insulator state and giant gauge factor of La0.7Sr0.3CoO3 films. Phys. Rev. Lett. 2008, 100, 076401. [Google Scholar] [CrossRef]
  100. Lee, Y.L.; Kleis, J.; Rossmeisl, J.; Shao-Horn, Y.; Morgan, D. Prediction of solid oxide fuel cell cathode activity with first-principles descriptors. Energy Environ. Sci. 2011, 4, 3966–3970. [Google Scholar] [CrossRef]
  101. Zhu, Z.; Shi, Y.; Aruta, C.; Yang, N. Improving electronic conductivity and oxygen reduction activity in Sr-doped lanthanum cobaltite thin films: Cobalt valence state and electronic band structure effects. ACS Appl. Energy Mater. 2018, 1, 5308–5317. [Google Scholar] [CrossRef]
  102. Yamazoe, N.; Teraoka, Y. Oxidation catalysis of perovskites --- relationships to bulk structure and composition (valency, defect, etc.). Catal. Today 1990, 8, 175–199. [Google Scholar] [CrossRef]
  103. Fierro, J.L.G.; Tejuca, L.G. Non-stoichiometric surface behaviour of LaMO3 oxides as evidenced by XPS. Appl. Surf. Sci. 1987, 27, 453–457. [Google Scholar] [CrossRef]
  104. Peña, M.A.; Fierro, J.L.G. Chemical structures and performance of perovskite oxides. Chem. Rev. 2001, 101, 1981–2018. [Google Scholar] [CrossRef]
  105. Chavan, S.V.; Singh, R.N. Preparation, properties, and reactivity of lanthanum strontium ferrite as an intermediate temperature SOFC cathode. J. Mater. Sci. 2013, 48, 6597–6604. [Google Scholar] [CrossRef]
  106. Simner, S.P.; Bonnett, J.F.; Canfield, N.L.; Meinhardt, K.D.; Shelton, J.P.; Sprenkle, V.L.; Stevenson, J.W. Development of lanthanum ferrite SOFC cathodes. J. Power Sources 2003, 113, 1–10. [Google Scholar] [CrossRef]
  107. Petitjean, M.; Caboche, G.; Siebert, E.; Dessemond, L.; Dufour, L.C. (La0.8Sr0.2)(Mn1−yFey)O3±δ oxides for IT-SOFC cathode materials?: Electrical and ionic transport properties. J. Eur. Ceram. Soc. 2005, 25, 2651–2654. [Google Scholar] [CrossRef]
  108. La O’, G.J.; Shao-Horn, Y. Oxygen surface exchange kinetics on Sr-substituted lanthanum manganite and ferrite thin-film microelectrodes. J. Electrochem. Soc. 2009, 156, B816–B824. [Google Scholar] [CrossRef]
  109. Wang, S.; Katsuki, M.; Dokiya, M.; Hashimoto, T. High temperature properties of La0.6Sr0.4Co0.8Fe0.2O3−δ phase structure and electrical conductivity. Solid State Ion. 2003, 159, 71–78. [Google Scholar] [CrossRef]
  110. Ralph, J.M.; Rossignol, C.; Kumar, R. Cathode materials for reduced-temperature SOFCs. J. Electrochem. Soc. 2003, 150, A1518–A1522. [Google Scholar] [CrossRef]
  111. Perry Murray, E.; Sever, M.J.; Barnett, S.A. Electrochemical performance of (La,Sr)(Co,Fe)O3–(Ce,Gd)O3 composite cathodes. Solid State Ion. 2002, 148, 27–34. [Google Scholar] [CrossRef]
  112. Adler, S.B.; Lane, J.A.; Steele, B.C.H. Electrode kinetics of porous mixed-conducting oxygen electrodes. J. Electrochem. Soc. 1996, 143, 3554–3564. [Google Scholar] [CrossRef]
  113. Chen, C.C.; Nasrallah, M.M.; Anderson, H.U.; Alim, M.A. Immittance response of La0.6Sr0.4Co0.2Fe0.8O3 based electrochemical cells. J. Electrochem. Soc. 1995, 142, 491–496. [Google Scholar] [CrossRef]
  114. Prestat, M.; Infortuna, A.; Korrodi, S.; Rey-Mermet, S.; Muralt, P.; Gauckler, L.J. Oxygen reduction at thin dense La0.52Sr0.48Co0.18Fe0.82O3–δ electrodes. J. Electroceram. 2007, 18, 111–120. [Google Scholar] [CrossRef]
  115. Angoua, B.F.; Slamovich, E.B. Single solution spray pyrolysis of La0.6Sr0.4Co0.2Fe0.8O3−δ–Ce0.8Gd0.2O1.9 (LSCF–CGO) thin film cathodes. Solid State Ion. 2012, 212, 10–17. [Google Scholar] [CrossRef]
  116. Develos-Bagarinao, K.; Kishimoto, H.; De Vero, J.; Yamaji, K.; Horita, T. Effect of La0.6Sr0.4Co0.2Fe0.8O3−δ microstructure on oxygen surface exchange kinetics. Solid State Ion. 2016, 288, 6–9. [Google Scholar] [CrossRef]
  117. Hopper, E.M.; Perret, E.; Ingram, B.J.; You, H.; Chang, K.C.; Baldo, P.M.; Fuoss, P.H.; Eastman, J.A. Oxygen exchange in La0.6Sr0.4Co0.2Fe0.8O3−δ thin-film heterostructures under applied electric potential. J. Phys. Chem. C 2015, 119, 19915–19921. [Google Scholar] [CrossRef]
  118. Plonczak, P.; Søgaard, M.; Bieberle-Hütter, A.; Hendriksen, P.V.; Gauckler, L.J. Electrochemical characterization of La0.58Sr0.4Co0.2Fe0.8O3−δ thin film electrodes prepared by pulsed laser deposition. J. Electrochem. Soc. 2012, 159, B471–B482. [Google Scholar] [CrossRef]
  119. Armstrong, E.N.; Duncan, K.L.; Oh, D.J.; Weaver, J.F.; Wachsman, E.D. Determination of surface exchange coefficients of LSM, LSCF, YSZ, GDC constituent materials in composite SOFC cathodes. J. Electrochem. Soc. 2011, 158, B492–B499. [Google Scholar] [CrossRef]
  120. Dumaisnil, K.; Carru, J.C.; Fasquelle, D.; Mascot, M.; Rolle, A.; Vannier, R.N. Promising performances for a La0.6Sr0.4Co0.8Fe0.2O3−δ cathode with a dense interfacial layer at the electrode-electrolyte interface. Ionics 2017, 23, 2125–2132. [Google Scholar] [CrossRef]
  121. Jang, I.; Kim, C.; Kim, S.; Yoon, H.; Paik, U. Fabrication of nanoparticle networked La0.6Sr0.4Co0.2Fe0.8O3−δ thin film layer between the cathode and electrolyte of solid oxide fuel cell by using a spin coating method. ECS Trans. 2017, 78, 741–745. [Google Scholar] [CrossRef]
  122. Jang, I.; Kim, S.; Kim, C.; Yoon, H.; Song, T. Enhancement of oxygen reduction reaction through coating a nano-web-structured La0.6Sr0.4Co0.2Fe0.8O3−δ thin-film as a cathode/electrolyte interfacial layer for lowering the operating temperature of solid oxide fuel cells. J. Power Sources 2018, 392, 123–128. [Google Scholar] [CrossRef]
  123. Ruddlesden, S.N.; Popper, P. The compound Sr3Ti2O7 and its structure. Acta Crystallogr. 1958, 11, 54–55. [Google Scholar] [CrossRef]
  124. Brown, I.D. Modeling the structures of La2NiO4. Z. Kristall. 1992, 199, 255–272. [Google Scholar] [CrossRef]
  125. Sase, M.; Yashiro, K.; Sato, K.; Mizusaki, J.; Kawada, T.; Sakai, N.; Yamaji, K.; Horita, T.; Yokokawa, H. Enhancement of oxygen exchange at the hetero interface of (La,Sr)CoO3/(La,Sr)2CoO4 in composite ceramics. Solid State Ion. 2008, 178, 1843–1852. [Google Scholar] [CrossRef]
  126. Lee, Y.L.; Lee, D.; Wang, X.R.; Lee, H.N.; Morgan, D.; Shao-Horn, Y. Kinetics of oxygen surface exchange on epitaxial Ruddlesden-Popper phases and correlations to first-principles descriptors. J. Phys. Chem. Lett. 2016, 7, 244–249. [Google Scholar] [CrossRef]
  127. Gopalakrishnan, J.; Colsmann, G.; Reuter, B. Studies on La2−xSrxNiO4 (0 <x <1) system. J. Solid State Chem. 1977, 22, 145–149. [Google Scholar]
  128. Bassat, J.M.; Burriel, M.; Wahyudi, O.; Castaing, R.; Ceretti, M.; Veber, P.; Weill, I.; Villesuzanne, A.; Grenier, J.C.; Paulus, W.; et al. Anisotropic oxygen diffusion properties in Pr2NiO4+δ and Nd2NiO4+δ single crystals. J. Phys. Chem. C 2013, 117, 26466–26472. [Google Scholar] [CrossRef]
  129. Burriel, M.; Santiso, J.; Rossell, M.D.; Van Tendeloo, G.; Figueras, A.; Garcia, G. Enhancing total conductivity of La2NiO4+δ epitaxial thin films by reducing thickness. J. Phys. Chem. C. 2008, 112, 10982–10987. [Google Scholar] [CrossRef]
  130. Garcia, G.; Burriel, M.; Bonanos, N.; Santiso, J. Electrical conductivity and oxygen exchange kinetics of La2NiO4+δ thin films grown by chemical vapor deposition. J. Electrochem. Soc. 2008, 155, P28–P32. [Google Scholar] [CrossRef]
  131. Kharton, V.V.; Yaremchenko, A.A.; Shaula, A.L.; Patrakeev, M.V.; Naumovich, E.N.; Loginovich, D.I.; Frade, J.R.; Marques, F.M.B. Transport properties and stability of Ni-containing mixed conductors with perovskite- and K2NiF4-type structure. J. Solid State Chem. 2004, 177, 26–37. [Google Scholar] [CrossRef]
  132. Kim, G.T.; Wang, S.Y.; Jacobson, A.J.; Yuan, Z.; Chen, C.L. Impedance studies of dense polycrystalline thin films of La2NiO4+δ. J. Mater. Chem. 2007, 17, 1316–1320. [Google Scholar] [CrossRef]
  133. Burriel, M.; Téllez, H.; Chater, R.J.; Castaing, R.; Veber, P.; Zaghrioui, M.; Ishihara, T.; Kilner, J.A.; Bassat, J.M. Influence of crystal orientation and annealing on the oxygen diffusion and surface exchange of La2NiO4+δ. J. Phys. Chem. C. 2016, 120, 17927–17938. [Google Scholar] [CrossRef]
  134. Yamada, A.; Suzuki, Y.; Saka, K.; Uehara, M.; Mori, D.; Kanno, R.; Kiguchi, T.; Mauvy, F.; Grenier, J.C. Ruddlesden-Popper-type epitaxial film as oxygen electrode for solid-oxide fuel cells. Adv. Mater. 2008, 20, 4124. [Google Scholar] [CrossRef]
  135. Tsvetkov, N.; Lu, Q.; Chen, Y.; Yildiz, B. Accelerated oxygen exchange kinetics on Nd2NiO4+δ thin films with tensile strain along c-axis. ACS Nano 2015, 9, 1613–1621. [Google Scholar] [CrossRef] [PubMed]
  136. Kim, G.; Wang, S.; Jacobson, A.J.; Chen, C.L. Measurement of oxygen transport kinetics in epitaxial LaNiO4+δ thin films by electrical conductivity relaxation. Solid State Ion. 2006, 177, 1461–1467. [Google Scholar] [CrossRef]
  137. Xie, W.; Lee, Y.L.; Shao-Horn, Y.; Morgan, D. Oxygen point defect chemistry in Ruddlesden–Popper oxides (La1–xSrx)2MO4±δ (M = Co, Ni, Cu). J. Phys. Chem. Lett. 2016, 7, 1939–1944. [Google Scholar] [CrossRef] [PubMed]
  138. Goodenough, J.B.; Manthiram, A. Crystal-chemistry and superconductivity of the copper oxides. J. Solid State Chem. 1990, 88, 115–139. [Google Scholar] [CrossRef]
  139. Gozar, A.; Logvenov, G.; Kourkoutis, L.F.; Bollinger, A.T.; Giannuzzi, L.A.; Muller, D.A.; Bozovic, I. High-temperature interface superconductivity between metallic and insulating copper oxides. Nature 2008, 455, 782. [Google Scholar] [CrossRef]
  140. Hoppler, J.; Stahn, J.; Niedermayer, C.; Malik, V.K.; Bouyanfif, H.; Drew, A.J.; Rossle, M.; Buzdin, A.; Cristiani, G.; Habermeier, H.U.; et al. Giant superconductivity-induced modulation of the ferromagnetic magnetization in a cuprate-manganite superlattice. Nat. Mater. 2009, 8, 315–319. [Google Scholar] [CrossRef]
  141. Opila, E.J.; Tuller, H.L. Thermogravimetric analysis and defect models of the oxygen nonstoichiometry in La2–xSrxCuO4–y. J. Am. Ceram. Soc. 1994, 77, 2727–2737. [Google Scholar] [CrossRef]
  142. Meyer, T.L.; Jacobs, R.; Lee, D.; Jiang, L.; Freeland, J.W.; Sohn, C.; Egami, T.; Morgan, D.; Lee, H.N. Strain control of oxygen kinetics in the Ruddlesden-Popper oxide La1.85Sr0.15CuO4. Nat. Commun. 2018, 9, 92. [Google Scholar] [CrossRef] [PubMed]
  143. Yashiro, K.; Nakamura, T.; Sase, M.; Hermes, F.; Sato, K.; Kawada, T.; Mizusaki, J. Composite cathode of perovskite-related oxides, (La,Sr)CoO3−δ/(La,Sr)2CoO4–δ, for solid oxide fuel cells. Electrochem. Solid State Lett. 2009, 12, B135–B137. [Google Scholar] [CrossRef]
  144. Chen, Y.; Téllez, H.; Burriel, M.; Yang, F.; Tsvetkov, N.; Cai, Z.; McComb, D.W.; Kilner, J.A.; Yildiz, B. Segregated chemistry and structure on (001) and (100) Surfaces of (La1–xSrx)2CoO4 override the crystal anisotropy in oxygen exchange kinetics. Chem. Mater. 2015, 27, 5436–5450. [Google Scholar] [CrossRef]
  145. Chen, J.; Liang, F.L.; Chi, B.; Pu, J.; Jiang, S.P.; Jian, L. Palladium and ceria infiltrated La0.8Sr0.2Co0.5Fe0.5O3–δ cathodes of solid oxide fuel cells. J. Power Sources 2009, 194, 275–280. [Google Scholar] [CrossRef]
  146. Han, J.W.; Yildiz, B. Mechanism for enhanced oxygen reduction kinetics at the (La,Sr)CoO3–δ/(La,Sr)2CoO4+δ hetero-interface. Energy Environ. Sci. 2012, 5, 8598–8607. [Google Scholar] [CrossRef]
  147. Liu, M.F.; Ding, D.; Blinn, K.; Li, X.X.; Nie, L.F.; Liu, M. Enhanced performance of LSCF cathode through surface modification. Int. J. Hydrogen Energy 2012, 37, 8613–8620. [Google Scholar] [CrossRef]
  148. Crumlin, E.J.; Mutoro, E.; Ahn, S.J.; la O’, G.J.; Leonard, D.N.; Borisevich, A.; Biegalski, M.D.; Christen, H.M.; Shao-Horn, Y. Oxygen reduction kinetics enhancement on a heterostructured oxide surface for solid oxide fuel cells. J. Phys. Chem. Lett. 2010, 1, 3149–3155. [Google Scholar] [CrossRef]
  149. Atkinson, A.; Ramos, T. Chemically-induced stresses in ceramic oxygen ion-conducting membranes. Solid State Ion. 2000, 129, 259–269. [Google Scholar] [CrossRef]
  150. Zhao, C.; Liu, X.; Zhang, W.; Zheng, Y.; Li, Y.; Yu, B.; Wang, J.; Chen, J. Measurement of oxygen reduction/evolution kinetics enhanced (La,Sr)CoO3/(La,Sr)2CoO4 hetero-structure oxygen electrode in operating temperature for SOCs. Int. J. Hydrog. Energy 2018. [Google Scholar] [CrossRef]
  151. Feng, Z.; Yacoby, Y.; Hong, W.T.; Zhou, H.; Biegalski, M.D.; Christen, H.M.; Shao-Horn, Y. Revealing the atomic structure and strontium distribution in nanometer-thick La0.8Sr0.2CoO3–δ grown on (001)-oriented SrTiO3. Energy Environ. Sci. 2014, 7, 1166–1174. [Google Scholar] [CrossRef]
  152. Feng, Z.; Crumlin, E.J.; Hong, W.T.; Lee, D.; Mutoro, E.; Biegalski, M.D.; Zhou, H.; Bluhm, H.; Christen, H.M.; Shao-Horn, Y. In situ studies of the temperature-dependent surface structure and chemistry of single-crystalline (001)-oriented La0.8Sr0.2CoO3−δ perovskite thin films. J. Phys. Chem. Lett. 2013, 4, 1512–1518. [Google Scholar] [CrossRef] [PubMed]
  153. Lou, X.Y.; Wang, S.Z.; Liu, Z.; Yang, L.; Liu, M.L. Improving La0.6Sr0.4Co0.2Fe0.8O3−δ cathode performance by infiltration of a Sm0.5Sr0.5CoO3− δ coating. Solid State Ion. 2009, 180, 1285–1289. [Google Scholar] [CrossRef]
  154. Nie, L.F.; Liu, M.F.; Zhang, Y.J.; Liu, M.L. La0.6Sr0.4Co0.2Fe0.8O3−δ cathodes infiltrated with samarium-doped cerium oxide for solid oxide fuel cells. J. Power Sources 2010, 195, 4704–4708. [Google Scholar] [CrossRef]
  155. Chen, H.; Guo, Z.; Zhang, L.A.; Li, Y.; Li, F.; Zhang, Y.; Chen, Y.; Wang, X.; Yu, B.; Shi, J.M.; et al. Improving the electrocatalytic activity and durability of the La0.6Sr0.4Co0.2Fe0.8O3−δ cathode by surface modification. ACS Appl. Mater. Interfaces 2018, 10, 39785–39793. [Google Scholar] [CrossRef]
Figure 1. Schematic of possible elementary reaction steps during oxygen reduction reaction (ORR) and possible pathways for two different classes of cathode materials; (a) pure electronic conductor and (b) mixed ionic and electronic conducting (MIEC) cathodes.
Figure 1. Schematic of possible elementary reaction steps during oxygen reduction reaction (ORR) and possible pathways for two different classes of cathode materials; (a) pure electronic conductor and (b) mixed ionic and electronic conducting (MIEC) cathodes.
Applsci 09 01030 g001
Figure 2. Schematic crystal structure of ABO3 perovskite oxides, where A and B are a rare earth and transition metal atom.
Figure 2. Schematic crystal structure of ABO3 perovskite oxides, where A and B are a rare earth and transition metal atom.
Applsci 09 01030 g002
Figure 3. Arrhenius plots of surface exchange coefficients (k*) for epitaxial La1−xSrxCoO3−δ thin films. (a) Effect of thickness difference on k* for La0.6Sr0.4CoO3−δ [96] and of substrate difference (STO and LAO) on k* for La0.8Sr0.2CoO3−δ [64] thin films and La0.8Sr0.2CoO3−δ [9] polycrystalline samples. (b) Effect of Sr substitution on k* for La1−xSrxCoO3−δ (x = 0, 0.2 and 0.4, [64,95,96]) thin films and La0.8Sr0.2CoO3−δ [9] polycrystalline oxides.
Figure 3. Arrhenius plots of surface exchange coefficients (k*) for epitaxial La1−xSrxCoO3−δ thin films. (a) Effect of thickness difference on k* for La0.6Sr0.4CoO3−δ [96] and of substrate difference (STO and LAO) on k* for La0.8Sr0.2CoO3−δ [64] thin films and La0.8Sr0.2CoO3−δ [9] polycrystalline samples. (b) Effect of Sr substitution on k* for La1−xSrxCoO3−δ (x = 0, 0.2 and 0.4, [64,95,96]) thin films and La0.8Sr0.2CoO3−δ [9] polycrystalline oxides.
Applsci 09 01030 g003
Figure 4. (a) Schematic electronic structure plots to illustrate the correlation of the ORR energetics vs. the O p-band center based on the rigid band model. Reprinted from [100] with permission of The Royal Society of Chemistry. (b) DFT calculations for the oxygen 2p-band center with respect to the Fermi level as a function of strain in LSCO films. The calculated oxygen 2p-band centers are averaged over LSCO films with five different Sr orderings. Reprinted from [96] with permission of The American Chemical Society.
Figure 4. (a) Schematic electronic structure plots to illustrate the correlation of the ORR energetics vs. the O p-band center based on the rigid band model. Reprinted from [100] with permission of The Royal Society of Chemistry. (b) DFT calculations for the oxygen 2p-band center with respect to the Fermi level as a function of strain in LSCO films. The calculated oxygen 2p-band centers are averaged over LSCO films with five different Sr orderings. Reprinted from [96] with permission of The American Chemical Society.
Applsci 09 01030 g004
Figure 5. Arrhenius plots of surface exchange coefficients (k*) for La1−xSrxCo1−yFeyO3−δ thin films. (a) Effect of strain on k* for La0.6Sr0.4Co0.2Fe0.8O3−δ [117] and La0.6Sr0.4Co0.2Fe0.8O3−δ [118] films and La0.6Sr0.4Co0.2Fe0.8O3−δ (Ref. [119]) polycrystalline samples. (b) Effect of Sr substitution on k* for La1−xSrxCo1−yFeyO3−δ (red [117], black [47] and blue [114]) films and La0.6Sr0.4Co0.2Fe0.8O3−δ [119] polycrystalline oxides.
Figure 5. Arrhenius plots of surface exchange coefficients (k*) for La1−xSrxCo1−yFeyO3−δ thin films. (a) Effect of strain on k* for La0.6Sr0.4Co0.2Fe0.8O3−δ [117] and La0.6Sr0.4Co0.2Fe0.8O3−δ [118] films and La0.6Sr0.4Co0.2Fe0.8O3−δ (Ref. [119]) polycrystalline samples. (b) Effect of Sr substitution on k* for La1−xSrxCo1−yFeyO3−δ (red [117], black [47] and blue [114]) films and La0.6Sr0.4Co0.2Fe0.8O3−δ [119] polycrystalline oxides.
Applsci 09 01030 g005
Figure 6. Schematic crystal structure of RP oxides, where A and B are a rare earth and transition metal atom.
Figure 6. Schematic crystal structure of RP oxides, where A and B are a rare earth and transition metal atom.
Applsci 09 01030 g006
Figure 7. Arrhenius plots of surface exchange coefficients (k*) for A2NiO4+δ (A = La and Nd): (a) Tensile (Tens.) and compressive (Comp.) strained (001)-oriented La2NiO4+δ films [61] and tensile, zero, and compressive strained (100)-oriented Nd2NiO4+δ thin films [135]; (b) (001)- and (100)-oriented La2NiO4+δ [61,63] thin films with various thickness and substrates.
Figure 7. Arrhenius plots of surface exchange coefficients (k*) for A2NiO4+δ (A = La and Nd): (a) Tensile (Tens.) and compressive (Comp.) strained (001)-oriented La2NiO4+δ films [61] and tensile, zero, and compressive strained (100)-oriented Nd2NiO4+δ thin films [135]; (b) (001)- and (100)-oriented La2NiO4+δ [61,63] thin films with various thickness and substrates.
Applsci 09 01030 g007
Figure 8. Oxygen surface exchange coefficients (k*) of La2CuO4 (light blue), LaSrCoO4 (green), La1.85Sr0.15CuO4 (blue), La1.6Sr0.4CuO4 (deep blue), and LaSrNiO4 (red) films with (001) orientations measured at 550 °C versus the computed DFT bulk O 2p-band centers (relative to the Fermi level, EFermi). The gray dashed lines represent the correlation between experimental oxygen surface-exchange coefficients and the computed bulk O 2p-band centers for a series of SOFC perovskites. Reprinted from [126] with permission of The American Chemical Society.
Figure 8. Oxygen surface exchange coefficients (k*) of La2CuO4 (light blue), LaSrCoO4 (green), La1.85Sr0.15CuO4 (blue), La1.6Sr0.4CuO4 (deep blue), and LaSrNiO4 (red) films with (001) orientations measured at 550 °C versus the computed DFT bulk O 2p-band centers (relative to the Fermi level, EFermi). The gray dashed lines represent the correlation between experimental oxygen surface-exchange coefficients and the computed bulk O 2p-band centers for a series of SOFC perovskites. Reprinted from [126] with permission of The American Chemical Society.
Applsci 09 01030 g008
Figure 9. Arrhenius plots of surface exchange coefficients (k*) for different RP oxides: (a) k* along c-direction in La2-xSrxMO4 (M = Cu, Co, Ni, [126,133,135,144]); and (b) k* along a-b plane in A2-xSrxMO4 (A = Sr, La, Nd and M = Co, Ni, [61,65,128,133,135,144]) thin films.
Figure 9. Arrhenius plots of surface exchange coefficients (k*) for different RP oxides: (a) k* along c-direction in La2-xSrxMO4 (M = Cu, Co, Ni, [126,133,135,144]); and (b) k* along a-b plane in A2-xSrxMO4 (A = Sr, La, Nd and M = Co, Ni, [61,65,128,133,135,144]) thin films.
Applsci 09 01030 g009
Figure 10. (a) The calculated Sr/La substitution energies in the bulk LSC214, LSC113, and LSCF113 (all relative to that of LSC113, which is set to 0). (b) The heterostructured interface model used in the DFT simulations and the results of Sr/La substitution energies. The elements are represented as: La/Sr (dark blue), Fe/Co (light blue, center of the octahedra), and O (red). Reprinted from [86] with permission of The Royal Society of Chemistry.
Figure 10. (a) The calculated Sr/La substitution energies in the bulk LSC214, LSC113, and LSCF113 (all relative to that of LSC113, which is set to 0). (b) The heterostructured interface model used in the DFT simulations and the results of Sr/La substitution energies. The elements are represented as: La/Sr (dark blue), Fe/Co (light blue, center of the octahedra), and O (red). Reprinted from [86] with permission of The Royal Society of Chemistry.
Applsci 09 01030 g010
Figure 11. Arrhenius plots of surface exchange coefficients (k*) for various heterostructure oxide thin films. Effect of heterostructure interfaces on k* for La0.6Sr0.4CoO3−δ/(LaSr)2CoO4±δ [150], La0.625Sr0.375Fe0.75Co0.25O3−δ/(LaSr)2CoO4±δ [86], La0.8Sr0.2CoO3−δ/La0.8Sr0.2MnO3−δ [76], La0.8Sr0.2CoO3−δ/(LaSr)2CoO4±δ [148] and La0.6Sr0.4Fe0.8Co0.2O3−δ/La0.4Sr0.6CoO3−δ:(LaSr)2CoO4±δ [66].
Figure 11. Arrhenius plots of surface exchange coefficients (k*) for various heterostructure oxide thin films. Effect of heterostructure interfaces on k* for La0.6Sr0.4CoO3−δ/(LaSr)2CoO4±δ [150], La0.625Sr0.375Fe0.75Co0.25O3−δ/(LaSr)2CoO4±δ [86], La0.8Sr0.2CoO3−δ/La0.8Sr0.2MnO3−δ [76], La0.8Sr0.2CoO3−δ/(LaSr)2CoO4±δ [148] and La0.6Sr0.4Fe0.8Co0.2O3−δ/La0.4Sr0.6CoO3−δ:(LaSr)2CoO4±δ [66].
Applsci 09 01030 g011

Share and Cite

MDPI and ACS Style

Yang, G.; Jung, W.; Ahn, S.-J.; Lee, D. Controlling the Oxygen Electrocatalysis on Perovskite and Layered Oxide Thin Films for Solid Oxide Fuel Cell Cathodes. Appl. Sci. 2019, 9, 1030. https://doi.org/10.3390/app9051030

AMA Style

Yang G, Jung W, Ahn S-J, Lee D. Controlling the Oxygen Electrocatalysis on Perovskite and Layered Oxide Thin Films for Solid Oxide Fuel Cell Cathodes. Applied Sciences. 2019; 9(5):1030. https://doi.org/10.3390/app9051030

Chicago/Turabian Style

Yang, Gene, Wonsang Jung, Sung-Jin Ahn, and Dongkyu Lee. 2019. "Controlling the Oxygen Electrocatalysis on Perovskite and Layered Oxide Thin Films for Solid Oxide Fuel Cell Cathodes" Applied Sciences 9, no. 5: 1030. https://doi.org/10.3390/app9051030

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop