Next Article in Journal
Attempt to Optimize the Corrosion Resistance of HRB400 Steel Rebar with Cr and RE
Next Article in Special Issue
Preparation of B4Cp/Al Composites via Selective Laser Melting and Their Tribological Properties
Previous Article in Journal
Study on the Magnetic Noise Characteristics of Amorphous and Nanocrystalline Inner Magnetic Shield Layers of SERF Co-Magnetometer
Previous Article in Special Issue
Microstructure and Properties of Densified Gd2O3 Bulk
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tailoring the Stability of Ti-Doped Sr2Fe1.4TixMo0.6−xO6−δ Electrode Materials for Solid Oxide Fuel Cells

1
Department of Hydrogen Energy, Faculty of Energy and Fuels, AGH University of Science and Technology, al. A. Mickiewicza 30, 30-059 Krakow, Poland
2
AGH Centre of Energy, AGH University of Science and Technology, ul. Czarnowiejska 36, 30-054 Krakow, Poland
3
Decentralised Hydrogen–Maciej Albrycht, ul. Wały Dwernickiego 21/23a, lok. 8, 42-200 Częstochowa, Poland
4
School of Materials Science and Energy Engineering, Foshan University, Foshan 528000, China
5
College of Pharmacy, Dali University, Dali 671000, China
6
Institute of Metallurgy and Materials Science, Polish Academy of Sciences, 25 Reymonta Str., 30-059 Krakow, Poland
*
Author to whom correspondence should be addressed.
Materials 2022, 15(22), 8268; https://doi.org/10.3390/ma15228268
Submission received: 22 October 2022 / Revised: 3 November 2022 / Accepted: 16 November 2022 / Published: 21 November 2022

Abstract

:
In this work, the stability of Sr2(FeMo)O6−δ-type perovskites was tailored by the substitution of Mo with Ti. Redox stable Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) perovskites were successfully obtained and evaluated as potential electrode materials for SOFCs. The crystal structure as a function of temperature, microstructure, redox stability, and thermal expansion properties in reducing and oxidizing atmospheres, oxygen content change, and transport properties in air and reducing conditions, as well as chemical stability and compatibility towards typical electrolytes have been systematically studied. All Sr2Fe1.4TixMo0.6−xO6−δ compounds exhibit a regular crystal structure with Pm-3m space group, showing excellent stability in oxidizing and reducing conditions. The increase of Ti-doping content in materials increases the thermal expansion coefficient (TEC), oxygen content change, and electrical conductivity in air, while it decreases the conductivity in reducing condition. All three materials are stable and compatible with studied electrolytes. Interestingly, redox stable Sr2Fe1.4Ti0.1Mo0.5O6−δ, possessing 1 μm grain size, low TEC (15.3 × 10−6 K−1), large oxygen content change of 0.72 mol·mol−1 between 30 and 900 °C, satisfactory conductivity of 4.1–7.3 S·cm−1 in 5% H2 at 600–800 °C, and good transport coefficients D and k, could be considered as a potential anode material for SOFCs, and are thus of great interest for further studies.

1. Introduction

Solid oxide fuel cells (SOFCs) with the advantage of high efficiency and fuel flexibility are among the most promising devices for the generation of electrical energy and heat from renewable and traditional energy sources, considerably reducing the emission of CO2 and other harmful gases (NOx, SOx, CO) [1,2,3,4,5] However, the high operational temperature (above 800 °C) of SOFCs leads to a high operational cost, limiting the choice of materials and delaying commercial applications of SOFCs. For practical use, SOFCs need to operate at much lower temperatures than the current range (≤800 °C) and still be able to generate a high-power production. Therefore, new anode [6,7,8] and cathode materials [9,10] with high stability and electrocatalytic activity are required to maintain a reasonable power output at temperature below 800 °C [11,12,13,14].
One group of the most interesting cathode and anode material candidates for SOFCs is the Sr2(FeMo)O6-type perovskite with Fe- and Mo-cations at B-site [15,16,17,18,19,20,21]. The B-site rock salt-type ordered Sr2−xBaxM1−yMoyO6 (M = Mg, Mn, Fe, Co, Ni) double perovskite-type compounds were initially proposed and evaluated as novel anode materials for SOFCs showing promising cell performance in different fuels, including hydrogen [22,23,24,25,26] and methane [27,28]. Generally, the choice of chemical composition in Sr2(Fe,Mo)O6−δ-type materials is governed by several crucial factors. The double perovskite (cations ordering) structure which favors the oxygen ion transport can be ensured by the considerable difference in oxidation state of B-site cations between Mo6+ or Mo5+ (in reducing atmospheres) and larger M2+/3+ cations (3d elements and Mg) [6,29]. The redox couple of M2+/M3+ and Mo6+/Mo5+ present in Sr2(Fe,Mo)O6−δ-type materials can not only favour an effective charge transport providing excellent conductivity [24,26], but also facilitate the creation of oxygen vacancies. For instance, Sr2FeMoO6−δ oxide shows very high metallic conductivity with 1000 S cm−1 in reducing condition, while unfortunately the compound is not stable in air at high temperatures [26]. The modification of B-site cations in Sr2(Fe,Mo)O6−δ-type double perovskites can possibly bring good redox stability in both oxidizing and reducing conditions [30,31,32]. SrFe0.75Mo0.25O3−δ [30,31], SrFe0.5Mn0.25Mo0.25O3−δ [30], Sr1−xBaxFe0.75W0.25O3−δ [33], and Sr2Fe1.2Mg0.2Mo0.6O6−δ and Sr2Fe0.9Mg0.4Mo0.7O6−δ perovskites [34] show good redox stability both in air and reducing conditions and have been studied as both cathode and anode materials for SOFCs. However, Sr2Fe1.5Mo0.5O6−δ compound is sensitive to water, and the reaction of Sr2Fe1.5Mo0.5O6−δ with H2O is a possible shortcoming of this perovskite as cathode and anode materials for SOFCs [35,36]. Mg-doped Sr2FeMo2/3Mg1/3O6−δ double perovskite with a good tolerance to sulfur poisoning and carbon deposition was evaluated as a promising anode material candidate for SOFCs [37]. The FeB-O-FeB′ bonds in materials promote easy creation of oxygen vacancies and their fast migration. However, the low electrical conductivity (4–5 S cm−1 at 600–800 °C in air) may limit the application of such a material for SOFCs [37]. The copper-substituted Sr2Fe1.5Mo0.3Cu0.2O6−δ material was investigated as a fuel electrode for the oxidation of H2 and CO2-CO reduction, showing improved reaction activity and durability, with an excellent SOFC power yield of 1.51 W cm−2 in H2 and a very good current density in the reduction of CO2 (with 1.94 A cm−2 at 1.4 V) [38]. The effect of Co-doping on the physicochemical and electrochemical properties of SrFe0.45Co0.45Mo0.1O3−δ double perovskite has been investigated [39]. The material has been proposed as an air electrode for reversible solid oxide fuel cells, with high mobility of electron holes and oxygen ions. However, the recorded high thermal expansion coefficient is a limiting issue for the potential application of Co-doped oxide [39]. Sr2Mg1−xCoxMoO6−δ perovskites with Co-doping at Mg-site, were investigated as novel anode materials for SOFCs, showing small anode polarization resistances [40], and the cobalt-doping positively contributes to the sinterability and ionic conductivity of materials [40]. However, the relatively weak bonding between Co-O is ascribed to the instability issue in anode condition, causing the reduction of Co to metallic cobalt. In addition, (PrBa)0.95(Fe0.9Mo0.1)2O5+δ double perovskite with very high conductivity (59.2 S cm−1 in 5% H2, and 217 S cm−1 in air at 800 °C) was evaluated as an anode candidate for SOFCs, presenting excellent cell performance with 1.18 W cm−2 at 800 °C in H2S-containing fuel [8,11]. Therefore, the further modification of Sr2(FeMo)O6-type perovskites will contribute to the development of novel cathode and anode materials with high performance for SOFCs.
SrTiO3-derived perovskites are also of great interest as novel electrode materials for SOFCs, characterized by high redox stability in oxidizing and reducing conditions, and by a low economic and environmental impact [41,42,43]. Perovskites with mixed Fe- and Ti- at B-sites can favour the oxygen reduction reaction for SOFCs [44]. Ti- and Mo-containing Sr2TiMoO6−δ double perovskite oxide was studied as novel anode material for SOFC, and it demonstrates excellent stability, sulphur poisoning resistance, and coking tolerance, as well as a good power output of 275 mW cm−2 in H2S-containing syngas at 850 °C [45]. BaxSr1−xTi1−yMoyO3 materials were synthesised and investigated as potential catalysts for the oxidation of CO and methane reforming [46]. In addition, the electrical conductivity and sintering properties of Ti-containing La0.5Sr1.5Ti1.5M0.5O6−δ (M = Fe, Co and Ni) double perovskites were evaluated in terms of their application as new anode material candidates for [47]. Sr2ScTi1−xMoxO6 (x = 0.1 and 0.5) double perovskites were studied as both cathode and anode materials for symmetrical solid oxide fuel cells, exhibiting a good power output of 218 mW cm−2 at 800 °C in humidified CH4 [48]. Ti-containing Sr2TiNi0.5Mo0.5O6−δ perovskite was also proposed and systematically evaluated as a new anode material for SOFCs, generating a power density of 335 mW cm−2 at 800 °C in humidified H2 [49].
It has been reported that the electrocatalytic activity In the series of Sr2Fe2−xMoxO6−δ double perovskites enhances with the increase of Mo-doping content, which contributes to the improved performance of SOFCs cells in different fuels (H2 and methanol) [32], and SOFC with Sr2Fe1.4Mo0.6O6−δ anode material shows much better electrochemical performance than the SOFC cell with redox stable Sr2Fe1.5Mo0.5O6−δ electrode [32]. Sr2Fe1.4Mo0.6O6−δ double perovskite possess very high electrical conductivity with 330 S cm−1 at 500 °C, nonetheless unfortunately it decomposes in air above 500 °C [50]. Interestingly, the Ti-doping at Fe-site in Sr2Fe1.4−xTixMo0.6O6−δ double perovskites significantly improves the structural stability of materials, which were systematically investigated as new anode material candidates for SOFCs [50]. Sr2Fe1.3Ti0.1Mo0.6O6−δ anode-based SOFC cell delivers a promising power output with >0.64 W cm2 in wet hydrogen at 900 °C [50]. However, such a material is still not stable in air at a high temperature range. Therefore, in this work, the stability of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) perovskites was tailored by the substitution of Mo with Ti at B-site. Sr2Fe1.4TixMo0.6−xO6−δ materials with excellent stability in both reducing and oxidizing atmospheres have been obtained. The crystal structure as a function of temperature, microstructure, redox stability, and thermal expansion properties in reducing and oxidizing atmospheres, oxygen content change and transport properties in air and reducing condition, as well as the chemical stability and compatibility towards typical solid electrolytes have been evaluated for the studied materials in terms of their application as electrode material candidates for SOFCs.

2. Materials and Methods

The synthesis of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) materials was conducted using the high temperature method (solid state reaction), with the calculated amounts of SrCO3, Fe2O3, TiO2, and MoO3 (all compounds with ≥ 99.9% purity) chemicals. All required chemicals after milling in the high efficiency planetary ball mill (Spex Sample-Prep 8000 M, Spex Sampleprep, Metuchen, UK) were pressed into pellets and fired in air for 10 h at 1200 °C. The crystal structure properties of all obtained materials were investigated by the XRD measurements within 10–110 deg. 2 Theta range using the Panalytical Empyrean diffractometer (CuKα radiation, Malvern, UK).
The chemical stability of Sr2Fe1.4TixMo0.6−xO6−δ compounds was studied by the reduction of oxides in 5 vol.% H2 in argon for 10 h at 1200 °C. The XRD data refinement was applied applying the Rietveld method using GSAS/EXPGUI software [51,52]. Scanning electron microscopy (SEM) studies of reduced powders were conducted using FEI Nova NanoSEM 200 apparatus. The high temperature XRD studies were performed for the in-situ oxidation in air of reduced Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) samples using Panalytical Empyrean apparatus with Anton Paar HTK 1200N oven-chamber. The in-situ oxidation of Sr2Fe1.4TixMo0.6−xO6−δ sinters was also investigated with the thermal expansion measurements in air from room temperature to 900 °C using the Linseis L75 Platinum Series dilatometer (Selb, Germany).
Thermogravimetric (TG) studies were carried out on the TA Instruments Q5000IR apparatus (New Castle, DE, USA) and STA PT1600 TG with differential scanning calorimetry (DSC) studies from 30 to 900 °C in different conditions (in air and in 5 vol.% H2/Argon) with the rate of 2°·min−1. The buoyancy effect was also taken into account. The electrical conductivity (σ) of Sr2Fe1.4TixMo0.6−xO6−δ samples was recorded to 900 °C in air and 5 vol.% H2 in argon by a four-probe DC technique, on the dense cuboid shape sinters. The porosity effect of the studied sinters was also considered [53]. The oxygen diffusion coefficient D and surface exchange constant k of Sr2Fe1.4Ti0.1Mo0.5O6−δ compound were studied using the mass relaxation technique in TA Instruments Q5000 IR on a very thin-sheet shape sample [54,55]. The mass relaxation measurements were performed with a very fast oxygen partial pressure change between 0.21 atm and 0.01 atm. The determination of coefficients D and k was conducted in custom-made Matlab code, based on Crank’s mathematical solutions [56]. The chemical stability and compatibility studies of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) materials towards typical electrolytes, such as CGO20–Ce0.8Gd0.2O1.9 and LSGM–La0.8Sr0.2Ga0.8Mg0.2O3−d solid electrolytes, were investigated by examining the XRD data gathered for the respective oxide and solid electrolyte mixtures (with a 50:50 wt.%), fired in air for 10 h at 1200 °C.

3. Results and Discussion

3.1. Crystal Structure and Microstructure

The obtained XRD results with the Rietveld refinement in Figure 1 and Table 1 show that as-synthesized Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) samples exhibit simple perovskite crystal structure, belonging to the cubic Pm-3m space-group. However, the Sr2Fe1.4Ti0.1Mo0.5O6−δ material synthesized in air possesses about 3.6% secondary phase (SrMoO4), which can be successfully removed by annealing the compound in reducing condition (see Figure 2a). As evidenced, the substitution of molybdenum by titanium in Sr2Fe1.4TixMo0.6−xO6−δ perovskites does not change the crystal symmetry but leads to a decrease of the unit cell parameter a. The doping of Ti4+ (with smaller oxidation state) at Mo6+ site contributes to the increase of Fe4+ (reducing the amount of Fe3+) and the content of oxygen vacancies in materials. The larger difference in ionic radius between Fe3+ (rFe3+ = 0.645 Å) and Fe4+ (rFe4+ = 0.585 Å) causes the decrease of the unit cell parameter, despite Ti4+ (rTi4+ = 0.605 Å) presenting a slightly bigger ionic radius than Mo6+ (rMo6+ = 0.59 Å). The geometric tolerance factor tg of all studied materials was calculated using the following equation of t g = [ A O ] 2 [ B O ] , where [A − O] and [B − O] are the refined geometric average values of interatomic distances of Sr-O and M-O (M: Fe, Ti and Mo), respectively [26]. The geometric tolerance factor was calculated to be tg = 1.00 for all measured Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) samples, which indicates the presence of a regular crystal structure in the studied materials. Similar compositions, such as Sr2Fe1.5Mo0.5O6−δ and SrFe0.5Mn0.25Mo0.25O3−δ [30], Sr2TiFe0.5Mo0.5O6−δ [57], and SrFe0.45Co0.45Mo0.1O3−δ [39] oxides, also show a Pm-3m simple perovskite structure. Meanwhile, the Ti-doping at Fe-site in Sr2Fe1.4−xTixMo0.6O6−δ (x= 0 and 0.1) leads to a double perovskite structure with Fm-3m space group [50].
The redox stability of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) materials was studied by annealing the compounds in 5 vol.% H2 in Argon at 1200 °C for 10h. The collected XRD data (Figure 2) after the reduction measurements show the reduced Sr2Fe1.4TixMo0.6−xO6−δ still possess the same crystal structure (Pm-3m), and the reduced B-site cations (Fe, Ti and Mo cations) with larger ionic radius contribute to a larger unit cell parameter a and volume V (see Table 1). Interestingly, among all investigated Sr2Fe1.4TixMo0.6−xO6−δ materials, Sr2Fe1.4Ti0.1Mo0.5O6−δ oxide presents the smallest relative volume change ∆V = 0.52% between the reduced and oxidized samples. This indicates the possible application of such a compound in oxidizing and reducing conditions [54]. For comparison, the relative volume changes between the oxidized and reduced Sr2Fe1.5Mo0.5O6−δ and Sr2Fe0.9Mg0.4Mo0.7O6−δ oxides are larger and reach 1.18% [30] and 0.55% [34], respectively. The increase of Ti-doping in Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.2 and 0.3) leads to a significantly large volume change ∆V, which can contribute to larger thermal expansion coefficients (TEC) in reducing and oxidizing atmospheres.
SEM microphotographs of reduced Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) powders are shown in Figure 3. No substantial differences were observed in the microstructure for all studied materials, and the grain size is approximately 1 µm. Interestingly, the Ti-doping at Fe-site in Sr2Fe1.4−xTixMo0.6O6−δ materials (x = 0–0.2) introduces a totally different microstructure with well-sintered aggregates (primary grains/crystallites are not visible) reported in the work [50].

3.2. Redox Stability and Thermal Expansion Properties

The in-situ oxidation of reduced Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) samples was examined by the high temperature XRD (HT-XRD) measurements in air. As can be observed in Figure 4, the oxidation of the reduced materials starts around 250 °C and finishes below 400 °C, accompanied by the significant change of unit cell parameter a. This is associated with the oxidation of reduced B-site cations (Fe, Ti, and Mo). The introduction of a higher content of titanium in Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.2 and 0.3) leads to a more considerable unit cell parameter change. The initial linear increase of unit cell parameter recoded below 250 °C is attributed to the thermal expansion of reduced samples. Above 400 °C, the oxidized materials present a linear increase of unit cell parameter (indicating a linear thermal expansion behavior) to 850 °C in air, and the HT-XRD data allow to calculate the thermal expansion coefficient (TEC). The increase of Ti content in Sr2Fe1.4TixMo0.6−xO6−δ leads to an increase of TEC values. Sr2Fe1.4Ti0.1Mo0.5O6−δ sample shows the lowest TEC among all studied materials, with TEC = 17.4 × 10−6 K−1. While Sr2Fe1.4Ti0.3Mo0.3O6−δ presents a rather high TEC value of 22.1 × 10−6 K−1, which can be a shortcoming for applications as electrode materials for SOFCs. No phase transition was observed in the HT-XRD studies for all samples, and all Sr2Fe1.4TixMo0.6−xO6−δ materials are stable to 850 °C in air (Figure 5), possessing the same simple Pm-3m perovskite structure.
The in-situ oxidation of reduced Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) compounds was also investigated by dilatometry measurements in air (see Figure 6a). As in the case of HT-XRD measurements (Figure 4), the oxidation of reduced samples in dilatometry measurements (Figure 6a) occurs between 250 and 400 °C. A linear thermal expansion of reduced samples was recorded below 250 °C, and for oxidized materials a linear thermal expansion occurs in the range of 400–900 °C. Sr2Fe1.4Ti0.1Mo0.5O6−δ sinter shows the smallest TEC with 16.4 × 10−6 K−1 among all studied samples, and the value is comparable with the TEC measured by HT-XRD studies (Figure 6a). Data presented in Figure 6a show that the TEC value of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) increases with the increased content of Ti in materials.
In addition, the thermal expansion behavior of oxidized Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) was studied in air (Figure 6b). For all three samples, two slopes with an obvious bending at 400 °C can be observed. The slope in the range of 400–900 °C, corresponding to a higher TEC, is related with the chemical expansion, caused by the reduction of B-site cations and loss of lattice oxygen, which was observed in the TG studies (Figure 7a). Sr2Fe1.4Ti0.3Mo0.3O6−δ possesses the largest TEC value (22.1 × 10−6 K−1), while the TEC of Sr2Fe1.4Ti0.1Mo0.5O6−δ is the smallest (15.3 × 10−6 K−1), which is close to the TEC values of typical solid electrolytes for SOFCs, such as: La0.9Sr0.1Ga0.8Mg0.2O3−δ (TEC = 12.17 × 10−6 K−1), Zr0.85Y0.15O2−δ (TEC = 10.8 × 10−6 K−1), and Ce0.8Gd0.2O2−δ (TEC = 12.5 × 10−6 K−1) [26]. Moreover, the in-situ reduction of oxidized Sr2Fe1.4Ti0.1Mo0.5O6−δ sample was conducted in 5 vol. % H2/argon (Figure 6c), and a considerable nonlinear change due to the reduction of B-site cations occurs above 400 °C, which corresponds well with the significant mass reduction of Sr2Fe1.4Ti0.1Mo0.5O6−δ recorded in the TG measurement in Figure 7b. Meanwhile, a linear thermal expansion of reduced material was observed above 575 °C and it presents a relatively low TEC of 15.0 × 10−6 K−1. For the reduced Sr2Fe1.4Ti0.1Mo0.5O6−δ sample in 5 vol. % H2/argon, it exhibits a linear thermal expansion, with a small TEC value of 13.7 × 10−6 K−1, which favours the application of such a material in reducing conditions. Therefore, in the case of the application of redox stable Sr2Fe1.4Ti0.1Mo0.5O6−δ perovskite as electrode material for SOFCs, the TEC mismatch (causing delamination) problem is alleviated, possibly providing stable cell performance.
As can be observed in Table 2, the substitution of Mo by Ti at B-site in Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) materials leads to an increase of TEC values. Hence, the Ti-doping in Sr2Fe1.4TixMo0.6−xO6−δ does not benefit the thermal expansion properties and the titanium content in those materials should be restricted to a rather small level (such as x = 0.1).

3.3. Oxygen Content and Transport Properties

The oxygen content change of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) was recorded by TG measurements in air and in 5 vol.% H2 in argon (see Figure 7a,b), respectively. For the TG studies in air, a significant mass loss of all three samples occurs above 300 °C, indicating the generation of additional oxygen vacancies in the studied perovskites, according to the following reaction: O O X 1 / 2 O 2 + V O + 2 e , which corresponds well with the chemical expansion of materials observed in Figure 6b. The substitution of Mo6+ by Ti4+ in the studied Sr2Fe1.4TixMo0.6−xO6−δ materials leads to the increase of oxygen vacancies. The data presented in Figure 7a indicate oxygen content decrease of ca. 0.1 mol·mol−1 for x = 0.1 sample up to 900 °C, ca. 0.17 mol·mol−1 for x = 0.2, and the largest change of ca. 0.20 mol·mol−1 for x = 0.3 material, respectively. In the reducing condition (Figure 7b), more oxygen vacancies were created in the materials, related with the reduction of B-site cations (Fe, Ti, Mo) to lower oxygen states (Fe2+/Fe3+, Ti3+ and Mo5+/Mo4+). The largest oxygen content change of ca. 0.86 mol·mol−1 was recorded for Sr2Fe1.4Ti0.3Mo0.3O6−δ sample. In the case of Sr2Fe1.4Ti0.2Mo0.4O6−δ, the oxygen content decrease of ca. 0.84 mol·mol−1 was documented, while for Sr2Fe1.4Ti0.1Mo0.5O6−δ it presents a relatively smaller change of 0.72 mol·mol−1.
In the supplementary DSC studies (Figure 7c), no endothermic or exothermic effects were observed for the investigated compounds, confirming no phase transition recorded, which was also documented in the in-situ HT-XRD studies (see Figure 4).
The electrical conductivity σ data measured for Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides in air (Figure 8a) show a maximum value of σ with the increase of temperature: it increases firstly and then decreases. All three materials initially exhibit a linear relationship below 400 °C with quite similar activation energy (0.22–0.23 eV), indicating small polaron conduction behavior. Similar behavior was also observed for Sr2Fe1.2Mg0.2Mo0.6O6−δ and Sr2Fe0.9Mg0.4Mo0.7O6−δ perovskites [34]. In Sr2Fe1.4TixMo0.6−xO6−δ materials, the electrons are transmitted via Fe3+-O2-/Fe4+ network in air. The Ti4+ substation of Mo6+ at B-site leads to the increase of Fe4+ content, thus favouring the electrical conductivities in air. Sr2Fe1.4Ti0.3Mo0.3O6−δ sample shows the highest conductivity in air among all studied materials, with a peak value of 17.8 S·cm−1 at 500 °C. Meanwhile, in the case of Sr2Fe1.4Ti0.2Mo0.4O6−δ compound, the maximum conductivity value of 11.5 S·cm−1 was observed at around 600 °C. For Sr2Fe1.4Ti0.1Mo0.5O6−δ, the maximum value (9.4 S·cm−1) was documented at 700 °C. The electrical conductivity decreases with a further increase of temperature, which is related with the release of oxygen from the lattice ( O O X 1 / 2 O 2 + V O + 2 e ) at high temperature range breaking the Fe3+-O2—Fe4+ network, causing the decrease of electrical conductivity. Similar behaviour was also observed in SrFeO3-based materials [21,30].
In the atmosphere of 5 vol.% H2/Ar, all Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2, and 0.3) samples exhibit lower conductivity. In the reducing condition, the electrical conductivity of Fe-and Mo-containing perovskites is strongly related with the content of Fe2+/Fe3+–Mo6+/Mo5+ redox couples [24,26,30,38,50]. The Ti-doping in Sr2Fe1.4TixMo0.6−xO6−δ leads to the decrease of Fe2+/Fe3+–Mo6+/Mo5+ redox pairs, thus resulting in the decrease of electrical conductivity. Among all three investigated samples, Sr2Fe1.4Ti0.3Mo0.3O6−δ oxide possesses the lowest conductivity (1.2–2.9 in 5% H2/Ar 600–800 °C) with the largest activation energy (Ea = 0.39 eV). In the case of Sr2Fe1.4Ti0.1Mo0.5O6−δ perovskite, it shows relatively satisfactory conductivity value (4.1–7.3 S·cm−1 at 600–800 °C) with a small activation energy Ea = 0.25 eV. The measured electrical conductivity σ for Sr2Fe1.4Ti0.1Mo0.5O6−δ is higher than the conductivity values (see Table 3) of Sr2Fe1.5Mo0.3Cu0.2O6−δ [38], Sr2MgMoO6−δ [58], Sr2Fe0.9Mg0.4Mo0.7O6−δ [34], Sr2−xBaxMgMoO6−δ, and Sr2−xBaxMnMoO6−δ [24,26], but smaller than the values of Sr2Fe1.2Mg0.2Mo0.6O6−δ [34] and Sr2Fe1.3Ti0.1Mo0.6O6−δ [50].
For Sr2Fe1.4Ti0.1Mo0.5O6−δ sample, the chemical diffusion coefficient D and surface exchange constant k were evaluated by the mass relaxation study (see Figure 8b,c). The chemical diffusion coefficient D is in the range of 5.7 × 10−5 to 7.4 × 10−5 cm2 s−1 at 600–800 °C, with an activation energy Ea,D = 0.1 eV. While the surface exchange k is within the scope of 1.5 × 10−3–1.8 × 10−3 cm s−1, with a very small activation energy Ea,k = 0.08 eV. The determined chemical diffusion coefficient values are comparable with the D values measured for Sr2TiNi0.5Mo0.5O6−δ [49], Sr2Fe1.2Mg0.2Mo0.6O6−δ sample [34] and Sr2Fe1.4Mn0.1Mo0.5O6−δ [59] materials. While the k values of Sr2Fe1.4Ti0.1Mo0.5O6−δ are bigger than k values of Sr2Fe1.2Mg0.2Mo0.6O6−δ sample [34] and Sr2Fe1.4Mn0.1Mo0.5O6−δ [59]. The relatively good transport coefficients D and k evaluated for Sr2Fe1.4Ti0.1Mo0.5O6−δ compound indicate good ionic transport properties in such a material.
Table 3. The crystal structure, electrical conductivity, thermal expansion coefficient (TEC), redox stability and compatibility with electrolytes, as well as the application of Sr2(Fe,Mo)O6−δ-based compounds.
Table 3. The crystal structure, electrical conductivity, thermal expansion coefficient (TEC), redox stability and compatibility with electrolytes, as well as the application of Sr2(Fe,Mo)O6−δ-based compounds.
MaterialStructureElectrical Conductivity [S·cm−1]TEC Value [×10−6 K−1]StabilityTowards ElectrolyteApplicationRef.
Sr2Fe1.4Ti0.1Mo0.5O6−δPm-3m9.4 at 700 °C in air; 4.1–7.3 in 5% H2 at 600–800 °C15.3 in airredox stablestable with CGO and LSGMcathode and anode candidatethis work
Sr2Fe1.4Ti0.2Mo0.4O6−δPm-3m11.5 at 600 °C in air; 1.6–3.5 in 5% H2 at 600–800 °C19.5 in airredox stablestable with CGO and LSGMcathode and anode candidatethis work
Sr2Fe1.4Ti0.3Mo0.3O6−δPm-3m17.8 at 500 °C in air; 1.2–2.9 in 5% H2 at 600–800 °C22.1 in airredox stablestable with CGO and LSGMcathode and anode candidatethis work
Sr2Fe1.5Mo0.3Cu0.2O6−δFm-3m0.06–0.36 in 5% H2 at 600–850 °C-decomposed in H2-fuel electrode[38]
Sr2Fe1.5Mo0.5O6−δFm-3m or Pm-3m2.89–5.55 in 5% H2 at 600–850 °C; 13 in air at 400–600 °C13.5–18.3 in airredox stablestable with CGOcathode and anode candidate[30,38]
Sr2MgMoO6−δI-10.8 in 5% H2 at 800 °C; 0.003 at 800 °C in air-stable in 5%H2-anode candidate[58]
Sr2−xBaxMnMoO6−δP21/n and Fm-3m 0.24 to 1.41 in 5% H211.5 to 14.8 (x = 0) in airstable in 5% H2/Ar-anode candidate[24,26]
Sr2−xBaxMgMoO6−δI4/m and Fm-3m0.14 to 1.38 in 5% H213.8 to 18.2 (x = 0) in airredox stable-anode candidate[24,26]
Sr2Mg0.95Al0.05MoO6−δ-5.4 in 5% H2 at 800 °C-redox stableStable with LSGM and CGO, reacts with YSZanode candidate[60]
SrFe0.5Mn0.25Mo0.25O3−δPm-3m3 at 850 °C in air; 10 at 850 °C in 5% H212.9 to 14.5 in airredox stablestable with CGOcathode and anode candidate[30]
Sr2Fe1.2Mg0.2Mo0.6O6−δFm-3m56.2 to 42.7 at 600–800 °C in air; 7.9 to 10.3 at 600–800 °C in 5% H214.6 to 16.7 in 5%H2; 12.9 to 14.6 in airredox stablestable with CGO, reacts with LSGMcathode and anode candidate[34]
Sr2Fe0.9Mg0.4Mo0.7O6−δFm-3m7.9 to 7.5 at 600–800 °C in air; 0.3 at 600–800 °C in 5% H214.2 to 15.1 in 5%H2; 13.5 to 15.7 in airredox stablestable with CGO, reacts with LSGMcathode and anode candidate[34]
Sr2Fe1.5Mo0.4Nb0.1O6−δPnma30 in air at 550 °C16.1 in airstable in airstable with LSGMcathode candidate[61]
Sr1.9Fe1.5Mo0.3Cu0.2O6−δ-54.8 in air at 630 °C19.4 in airdecomposed in H2-anode candidate[62]
La0.5Sr0.5Fe0.9Mo0.1O3−δPm-3m2.7 to 6.7 at 600–800 °C in H215.1 in 5%H2; 13.4 in air stable <750 °C in H2stable with LSGMcathode and anode candidate[63]
Sr2FeMo2/3Mg1/3O6−δFm-3m4–5 in air at 600–800 °C; 9–13 in H2 at 600–800 °C16.9 in airredox stablestable with LDCanode candidate[37]
Sr2FeMo0.65Ni0.35O6−δI4/m55.4 in 5% H2 at 800 °C-decomposed in H2stable with LDC anode candidate[18]
Sr2Fe1.3Ti0.1Mo0.6O6−δFm-3m220 to 160 at 500–800 °C in 5% H213.5 at 550 °C in airstable in H2stable with CGOanode candidate[50]
Sr2TiFe0.5Mo0.5O6−δPm-3m22.3 in H2 at 800 °C11.2 in H2stable in H2stable with LSGM91 and CSOanode candidate[57]
Sr2TiNi0.5Mo0.5O6−δ-17.5 at 800 °C in hydrogen 12.8 in airstable in H2stable with LSGManode candidate[49]
Sr2−xBaxFeMoO6−δI4/m and Fm-3m100 to 1000 in 5% H213.8 (for x = 0) in airstable in 5%H2stable with CGOanode candidate[24,26]
SrFe0.45Co0.45Mo0.1O3−δPm-3m298 at 300 °C in air14.8 to 30.8 in airstable in air-air electrode
candidate
[39]
Sr2Mg0.3Co0.7MoO6−δI-19 to 7 at 600–800 °C in 5% H213.9 in air--anode candidate[40]
CGO: Ce0.8Gd0.2O1.9, LSGM: La0.8Sr0.2Ga0.8Mg0.2O3-d, YSZ: Zr0.92Y0.08O1.96, LDC: La0.4Ce0.6O2−d, LSGM91: La0.9Sr0.1Ga0.8Mg0.2O3−d, CSO: Ce0.8Sm0.2O1.9.

3.4. Chemical Stability and Compatibility with Electrolytes

The stability of cathode and anode oxides and their compatibility towards applied electrolytes at high temperature range are very critical for the electrode layer preparation (sintering) and a stable performance of cells. Therefore, the chemical stability of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides and their compatibility with typical solid electrolyte/buffer layer materials, such as Ce0.8Gd0.2O1.9–CGO20 and La0.8Sr0.2Ga0.8Mg0.2O3−d–LSGM, were evaluated by analyzing XRD data gathered for the respective materials and solid electrolyte powders (50:50 wt.%), which were fired in air for 10 h at 1200 °C. The Rietveld refined XRD patterns of Sr2Fe1.4TixMo0.6−xO6−δ with electrolyte (50:50 wt.%) powder mixtures are shown in Figure 9 and Figure 10. The XRD patterns present no chemical reaction occurring between the studied compounds and solid electrolytes (CGO20 and LSGM), confirming the good stability and compatibility of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) materials with used electrolytes.
As presented in Table 4, the unit cell parameter of Ce0.8Gd0.2O1.9 from different mixtures is very similar, and the unit cell parameters of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) from CGO20 mixtures are also very close to the values of appropriate materials recorded from LSGM mixtures. In addition, the unit cell parameter for LSGM from various mixtures is almost identical. This shows the good chemical stability of all investigated electrode materials and their compatibility towards electrolytes (Ce0.8Gd0.2O1.9 and La0.8Sr0.2Ga0.8Mg0.2O3−d). Interestingly, similar compositions Sr2Fe1.2Mg0.2Mo0.6O6−δ and Sr2Fe0.9Mg0.4Mo0.7O6−δ are compatible with Ce0.8Gd0.2O1.9 electrolyte but react with La0.8Sr0.2Ga0.8Mg0.2O3−d [34]. Sr2Fe1.3Ti0.1Mo0.6O6−δ (material unstable in air) is also compatible with Ce0.8Gd0.2O1.9 [50].
In addition, the long-term compatibility of Sr2Fe1.4Ti0.1Mo0.5O6−δ towards Ce0.8Gd0.2O1.9 and with La0.8Sr0.2Ga0.8Mg0.2O3−d solid electrolytes was evaluated, after firing for 100 h at 800 °C in air. As shown in Figure 11, Sr2Fe1.4Ti0.1Mo0.5O6−δ sample has no reaction with Ce0.8Gd0.2O1.9 and La0.8Sr0.2Ga0.8Mg0.2O3-d after annealing at 800 °C for 100 h. Moreover, the structural parameters gathered for Sr2Fe1.4Ti0.1Mo0.5O6−δ, Ce0.8Gd0.2O1.9 and La0.8Sr0.2Ga0.8Mg0.2O3−d materials in Table 5 are very close to the respective results in Table 4, which indicate that the Sr2Fe1.4Ti0.1Mo0.5O6−δ sample exhibits excellent stability and compatibility towards the used solid electrolytes, potentially providing a stable performance of Sr2Fe1.4Ti0.1Mo0.5O6−δ in the cell.

4. Summary

In this work, the stability of Sr2(FeMo)O6−δ-type perovskites was successfully tailored by the substitution of Mo with Ti at B-site, and Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) perovskites with excellent redox stability in reducing and oxidizing conditions were obtained. All Sr2Fe1.4TixMo0.6−xO6−δ materials possess a regular simple perovskite structure with Pm-3m space group, showing excellent stability in both reducing and oxidizing conditions up to 1200 °C. All three materials present a similar microstructure with 1 μm grain size. The in-situ oxidation of reduced samples, observed by HT-XRD measurements and dilatometry studies, shows that the increased content of Ti doping at Mo-site in materials increases the TEC values. Sr2Fe1.4Ti0.1Mo0.5O6−δ shows the lowest TEC with 15.3 × 10−6 K−1. In addition, Ti-doping also increases the oxygen content change and electrical conductivity in air, while it decreases the conductivity in reducing condition. Sr2Fe1.4Ti0.3Mo0.3O6−δ sample presents the highest conductivity in air with 17.8 S·cm−1 at 500 °C, while the high TEC value of 22.1 × 10−6 K−1 can potentially limit the application. All three Sr2Fe1.4TixMo0.6−xO6−δ materials are stable and compatible with studied electrolytes (Ce0.8Gd0.2O1.9 and La0.8Sr0.2Ga0.8Mg0.2O3−d).
Redox stable Sr2Fe1.4Ti0.1Mo0.5O6−δ seems to be the most interesting among studied materials, with large oxygen content change of 0.72 mol·mol−1 between 30 and 900 °C, satisfactory conductivity of 4.1–7.3 S·cm−1 in 5% H2 at 600–800 °C, and good transport coefficients D and k, which indicate that such a material can be considered as a potential anode material for SOFCs and is of great interest for further studies.

Author Contributions

Conceptualization, methodology, investigation, validation, formal analysis, resources, data curation, writing—original draft preparation, funding acquisition, supervision, K.Z.; visualization, investigation, M.A.; writing—review and editing, M.C., K.Q. and P.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Centre Poland (NCN) on the basis of the decision number UMO-2021/43/D/ST5/00824.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Boldrin, P.; Brandon, N.P. Progress and outlook for solid oxide fuel cells for transportation applications. Nat. Catal. 2019, 2, 571–577. [Google Scholar] [CrossRef] [Green Version]
  2. Park, S.; Vohs, J.M.; Gorte, R.J. Direct oxidation of hydrocarbons in a solid-oxide fuel cell. Nature 2000, 404, 265–267. [Google Scholar] [CrossRef] [PubMed]
  3. Develos-Bagarinao, K.; Ishiyama, T.; Kishimoto, H.; Shimada, H.; Yamaji, K. Nanoengineering of cathode layers for solid oxide fuel cells to achieve superior power densities. Nat. Commun. 2021, 12, 3979. [Google Scholar] [CrossRef]
  4. Irvine, J.T.S.; Neagu, D.; Verbraeken, M.C.; Chatzichristodoulou, C.; Graves, C.R.; Mogensen, M.B. Evolution of the electrochemical interface in high-temperature fuel cells and electrolysers. Nat. Energy 2016, 1, 15014. [Google Scholar] [CrossRef] [Green Version]
  5. Myung, J.-H.; Neagu, D.; Miller, D.N.; Irvine, J. Switching on electrocatalytic activity in solid oxide cells. Nature 2016, 537, 528–531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Huang, Y.-H.; Dass, R.I.; Xing, Z.-L.; Goodenough, J.B.; Goldman, A.S.; Roy, A.H.; Ahuja, R.; Schinski, W.; Brookhart, M. Double Perovskites as Anode Materials for Solid-Oxide Fuel Cells. Science 2006, 312, 254–257. [Google Scholar] [CrossRef]
  7. Tao, S.; Irvine, J. A redox-stable efficient anode for solid-oxide fuel cells. Nat. Mater. 2003, 2, 320–323. [Google Scholar] [CrossRef]
  8. Ding, H.; Tao, Z.; Liu, S.; Zhang, J. A High-Performing Sulfur-Tolerant and Redox-Stable Layered Perovskite Anode for Direct Hydrocarbon Solid Oxide Fuel Cells. Sci. Rep. 2015, 5, 18129. [Google Scholar] [CrossRef] [Green Version]
  9. Graves, C.R.; Ebbesen, S.D.; Jensen, S.H.; Simonsen, S.B.; Mogensen, M.B. Eliminating degradation in solid oxide electrochemical cells by reversible operation. Nat. Mater. 2015, 14, 239–244. [Google Scholar] [CrossRef]
  10. Hughes, G.A.; Railsback, J.G.; Yakal-Kremski, K.J.; Butts, D.M.; Barnett, S.A. Degradation of (La0.8Sr0.2)0.98MnO3−δ–Zr0.84Y0.16O2−γ composite electrodes during reversing current operation. Faraday Discuss. 2015, 182, 365–377. [Google Scholar] [CrossRef]
  11. Ding, H.; Fang, S.; Yang, Y.; Yang, Y.; Wu, W.; Tao, Z. High-performing and stable electricity generation by ceramic fuel cells operating in dry methane over 1000 hours. J. Power Sources 2018, 401, 322–328. [Google Scholar] [CrossRef]
  12. Laguna-Bercero, M.A. Recent advances in high temperature electrolysis using solid oxide fuel cells: A review. J. Power Sources 2012, 203, 4–16. [Google Scholar] [CrossRef] [Green Version]
  13. Hossain, M.K.; Raihan, G.A.; Akbar, M.A.; Rubel, M.H.K.; Ahmed, M.H.; Khan, M.I.; Hossain, S.; Sen, S.K.; Jalal, M.I.E.; El-Denglawey, A. Current Applications and Future Potential of Rare Earth Oxides in Sustainable Nuclear, Radiation, and Energy Devices: A Review. ACS Appl. Electron. Mater. 2022, 4, 3327–3353. [Google Scholar] [CrossRef]
  14. Hossain, M.K.; Chanda, R.; El-Denglawey, A.; Emrose, T.; Rahman, M.T.; Biswas, M.C.; Hashizume, K. Recent progress in barium zirconate proton conductors for electrochemical hydrogen device applications: A review. Ceram. Int. 2021, 47, 23725–23748. [Google Scholar] [CrossRef]
  15. Sengodan, S.; Choi, S.; Jun, A.; Shin, T.H.; Ju, Y.-W.; Jeong, H.Y.; Shin, J.; Irvine, J.; Kim, G. Layered oxygen-deficient double perovskite as an efficient and stable anode for direct hydrocarbon solid oxide fuel cells. Nat. Mater. 2015, 14, 205–209. [Google Scholar] [CrossRef] [PubMed]
  16. Skutina, L.; Filonova, E.; Medvedev, D.; Maignan, A. Undoped Sr2MMoO6 Double Perovskite Molybdates (M = Ni, Mg, Fe) as Promising Anode Materials for Solid Oxide Fuel Cells. Materials 2021, 14, 1715. [Google Scholar] [CrossRef] [PubMed]
  17. Zamudio-García, J.; Caizán-Juanarena, L.; Porras-Vázquez, J.M.; Losilla, E.R.; Marrero-López, D. A review on recent advances and trends in symmetrical electrodes for solid oxide cells. J. Power Sources 2022, 520, 230852. [Google Scholar] [CrossRef]
  18. Du, Z.; Zhao, H.; Yi, S.; Xia, Q.; Gong, Y.; Zhang, Y.; Cheng, X.; Li, Y.; Gu, L.; Świerczek, K. High-Performance Anode Material Sr2FeMo0.65Ni0.35O6−δ with In Situ Exsolved Nanoparticle Catalyst. ACS Nano 2016, 10, 8660–8669. [Google Scholar] [CrossRef]
  19. Zhu, K.; Luo, B.; Liu, Z.; Wen, X. Recent advances and prospects of symmetrical solid oxide fuel cells. Ceram. Int. 2022, 48, 8972–8986. [Google Scholar] [CrossRef]
  20. Zhang, B.; Wan, Y.; Hua, Z.; Tang, K.; Xia, C. Tungsten-Doped PrBaFe2O5+δ Double Perovskite as a High-Performance Electrode Material for Symmetrical Solid Oxide Fuel Cells. ACS Appl. Energy Mater. 2021, 4, 8401–8409. [Google Scholar] [CrossRef]
  21. Cao, Y.; Zhu, Z.; Zhao, Y.; Zhao, W.; Wei, Z.; Liu, T. Development of tungsten stabilized SrFe0.8W0.2O3−δ material as novel symmetrical electrode for solid oxide fuel cells. J. Power Sources 2020, 455, 227951. [Google Scholar] [CrossRef]
  22. Xiao, G.; Liu, Q.; Zhao, F.; Zhang, L.; Xia, C.; Chen, F. Sr2Fe1.5Mo0.5O6 as Cathodes for Intermediate -Temperature Solid Oxide Fuel Cells with La0.8Sr0.2Ga0.87Mg0.13O3 Electrolyte. J. Electrochem. Soc. 2011, 158, B455. [Google Scholar] [CrossRef] [Green Version]
  23. Zheng, K.; Świerczek, K.; Bratek, J.; Klimkowicz, A. Cation-ordered perovskite-type anode and cathode materials for solid oxide fuel cells. Solid State Ion. 2014, 262, 354–358. [Google Scholar] [CrossRef]
  24. Zheng, K.; Świerczek, K. A- and B-site doping effect on physicochemical properties of Sr2−xBaxMMoO6 (M == Mg, Mn, Fe) double perovskites—Candidate anode materials for SOFCs. Funct. Mater. Lett. 2016, 9, 1641002. [Google Scholar] [CrossRef]
  25. Zheng, K.; Świerczek, K.; Zając, W.; Klimkowicz, A. Rock salt ordered-type double perovskite anode materials for solid oxide fuel cells. Solid State Ion. 2014, 257, 9–16. [Google Scholar] [CrossRef]
  26. Zheng, K.; Świerczek, K. Physicochemical properties of rock salt-type ordered Sr2MMoO6 (M = Mg, Mn, Fe, Co, Ni) double perovskites. J. Eur. Ceram. Soc. 2014, 34, 4273–4284. [Google Scholar] [CrossRef]
  27. Huang, Y.-H.; Dass, R.I.; Denyszyn, J.C.; Goodenough, J.B. Synthesis and Characterization of Sr2MgMoO6−δ: An Anode Material for the Solid Oxide Fuel Cell. J. Electrochem. Soc. 2006, 153, A1266. [Google Scholar] [CrossRef]
  28. Marrero-López, D.; Peña-Martínez, J.; Ruiz-Morales, J.C.; Gabás, M.; Núñez, P.; Aranda, M.A.G.; Ramos-Barrado, J.R. Redox behaviour, chemical compatibility and electrochemical performance of Sr2MgMoO6−δ as SOFC anode. Solid State Ion. 2010, 180, 1672–1682. [Google Scholar] [CrossRef]
  29. Goodenough, J.B.; Huang, Y.-H. Alternative anode materials for solid oxide fuel cells. J. Power Sources 2007, 173, 1–10. [Google Scholar] [CrossRef]
  30. Zheng, K.; Świerczek, K.; Polfus, J.M.; Sunding, M.F.; Pishahang, M.; Norby, T. Carbon Deposition and Sulfur Poisoning in SrFe0.75Mo0.25O3−δ and SrFe0.5Mn0.25Mo0.25O3−δ Electrode Materials for Symmetrical SOFCs. J. Electrochem. Soc. 2015, 162, F1078–F1087. [Google Scholar] [CrossRef]
  31. Liu, Q.; Dong, X.; Xiao, G.; Zhao, F.; Chen, F. A Novel Electrode Material for Symmetrical SOFCs. Adv. Mater. 2010, 22, 5478–5482. [Google Scholar] [CrossRef] [PubMed]
  32. Li, H.; Zhao, Y.; Wang, Y.; Li, Y. Sr2Fe2−xMoxO6−δ perovskite as an anode in a solid oxide fuel cell: Effect of the substitution ratio. Catal. Today 2016, 259, 417–422. [Google Scholar] [CrossRef] [Green Version]
  33. Zheng, K.; Świerczek, K. Evaluation of W-containing Sr1 −xBaxFe0.75W0.25O3−δ (x = 0, 0.5, 1) anode materials for solid oxide fuel cells. Solid State Ion. 2016, 288, 124–129. [Google Scholar] [CrossRef]
  34. Zheng, K.; Lach, J.; Zhao, H.; Huang, X.; Qi, K. Magnesium-Doped Sr2(Fe,Mo)O6−δ Double Perovskites with Excellent Redox Stability as Stable Electrode Materials for Symmetrical Solid Oxide Fuel Cells. Membranes 2022, 12, 1006. [Google Scholar] [CrossRef]
  35. Wright, J.H.; Virkar, A.V.; Liu, Q.; Chen, F. Electrical characterization and water sensitivity of Sr2Fe1.5Mo0.5O6−δ as a possible solid oxide fuel cell electrode. J. Power Sources 2013, 237, 13–18. [Google Scholar] [CrossRef]
  36. Fang, T.-T.; Ko, T.-F. Factors Affecting the Preparation of Sr2Fe2−xMoxO6. J. Am. Ceram. Soc. 2003, 86, 1453–1455. [Google Scholar] [CrossRef]
  37. Du, Z.; Zhao, H.; Li, S.; Zhang, Y.; Chang, X.; Xia, Q.; Chen, N.; Gu, L.; Świerczek, K.; Li, Y.; et al. Exceptionally High Performance Anode Material Based on Lattice Structure Decorated Double Perovskite Sr2FeMo2/3Mg1/3O6−δ for Solid Oxide Fuel Cells. Adv. Energy Mater. 2018, 8, 1800062. [Google Scholar] [CrossRef]
  38. He, F.; Hou, M.; Zhu, F.; Liu, D.; Zhang, H.; Yu, F.; Zhou, Y.; Ding, Y.; Liu, M.; Chen, Y. Building Efficient and Durable Hetero-Interfaces on a Perovskite-Based Electrode for Electrochemical CO2 Reduction. Adv. Energy Mater. 2022, 2202175. [Google Scholar] [CrossRef]
  39. Zapata-Ramírez, V.; Mather, G.C.; Azcondo, M.T.; Amador, U.; Perez-Coll, D. Electrical and electrochemical properties of the Sr(Fe,Co,Mo)O3−δ system as air electrode for reversible solid oxide cells. J. Power Sources 2019, 437, 226895. [Google Scholar] [CrossRef]
  40. Xie, Z.; Zhao, H.; Du, Z.; Chen, T.; Chen, N.; Liu, X.; Skinner, S.J. Effects of Co Doping on the Electrochemical Performance of Double Perovskite Oxide Sr2MgMoO6−δ as an Anode Material for Solid Oxide Fuel Cells. J. Phys. Chem. C 2012, 116, 9734–9743. [Google Scholar] [CrossRef]
  41. Zhang, Y.; Yu, Z.; Tao, Y.; Lu, J.; Liu, Y.; Shao, J. Insight into the electrochemical processes of the titanate electrode with in situ Ni exsolution for solid oxide cells. ACS Appl. Energy Mater. 2019, 2, 4033–4044. [Google Scholar] [CrossRef]
  42. Miller, D.N.; Irvine, J.T.S. B-site doping of lanthanum strontium titanate for solid oxide fuel cell anodes. J. Power Sources 2011, 196, 7323–7327. [Google Scholar] [CrossRef]
  43. Neagu, D.; Irvine, J.T.S. Enhancing electronic conductivity in strontium titanates through correlated A and B-site doping. Chem. Mater. 2011, 23, 1607–1617. [Google Scholar] [CrossRef]
  44. Lan, C.; Luo, J.; Dou, M.; Zhao, S. First-principles calculations of the oxygen-diffusion mechanism in mixed Fe/Ti perovskites for solid-oxide fuel cells. Ceram. Int. 2019, 45, 17646–17652. [Google Scholar] [CrossRef]
  45. Niu, B.; Jin, F.; Liu, J.; Zhang, Y.; Jiang, P.; Feng, T.; Xu, B.; He, T. Highly carbone and sulfuretolerant Sr2TiMoO6−δ double perovskite anode for solid oxide fuel cellsInt. Int. J. Hydrogen Energy 2019, 44, 20404–20415. [Google Scholar] [CrossRef]
  46. Carollo, G.; Garbujo, A.; Mauvy, F.; Glisenti, A. Critical Raw Material-Free Catalysts and Electrocatalysts: Complementary Strategies to Activate Economic, Robust, and Ecofriendly SrTiO3. Energy Fuels 2020, 34, 11438–11448. [Google Scholar] [CrossRef]
  47. Ke, M.; Wang, W.; Yang, X.; Li, B.; Li, H. Doped Strontium Titanate Anode for Solid Oxide Fuel Cells: Electrical and Sintering Behavior. Ceram. Int. 2022, 48, 8709–8714. [Google Scholar] [CrossRef]
  48. Rath, M.K.; Kossenko, A.; Zinigrad, M.; Kalashnikov, A. In-operando gas switching to suppress the degradation of symmetrical solid oxide fuel cells. J. Power Sources 2020, 476, 228630. [Google Scholar] [CrossRef]
  49. He, B.; Wang, Z.; Zhao, L.; Pan, X.; Wu, X.; Xia, C. Ti-doped molybdenum-based perovskites as anodes for solid oxide fuel cells. J. Power Sources 2013, 241, 627–633. [Google Scholar] [CrossRef]
  50. Zheng, K. Ti-doped Sr2Fe1.4−xTixMo0.6O6−δ double perovskites with improved stability as anode materials for Solid Oxide Fuel Cells. Mater. Res. Bull. 2020, 128, 110877. [Google Scholar] [CrossRef]
  51. Larson, A.C.; Von Dreele, R.B. General Structure Analysis System (GSAS). Los Alamos National Laboratory Report LAUR 86-748; 2004. Available online: https://11bm.xray.aps.anl.gov/documents/GSASManual.pdf (accessed on 10 January 2022).
  52. Toby, B.H. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef] [Green Version]
  53. Stroud, D. Generalized effective-medium approach to the conductivity of an inhomogeneous material. Phys. Rev. B 1975, 12, 3368–3373. [Google Scholar] [CrossRef]
  54. Zheng, K. Enhanced oxygen mobility by doping Yb in BaGd1−xYbxMn2O5+δ double perovskite-structured oxygen storage materials. Solid State Ion. 2019, 335, 103–112. [Google Scholar] [CrossRef]
  55. Zheng, K.; Świerczek, K. Possibility of determination of transport coefficients D and k from relaxation experiments for sphere-shaped powder samples. Solid State Ion. 2018, 323, 157–165. [Google Scholar] [CrossRef]
  56. Crank, J. The Mathematics of Diffusion, 2nd ed.; Oxford University Press: New York, NY, USA, 1975. [Google Scholar]
  57. Niu, B.; Jin, F.; Yang, X.; Feng, T.; He, T. Resisting coking and sulfur poisoning of double perovskite Sr2TiFe0.5Mo0.5O6−δ anode material for solid oxide fuel cells. Int. J. Hydrogen Energy 2018, 43, 3280–3290. [Google Scholar] [CrossRef]
  58. Marrero-Lόpez, D.; Peña-Martínez, J.; Ruiz-Morales, J.C.; Perez-Coll, D.; Aranda, M.A.G.; Núñez, P. Synthesis, phase stability and electrical conductivity of Sr2MgMoO6−δ anode. Mater. Res. Bull. 2008, 43, 2441–2450. [Google Scholar] [CrossRef]
  59. Jiang, Y.; Yang, Y.; Xia, C.; Bouwmeester, H.J.M. Sr2Fe1.4Mn0.1Mo0.5O6−δ perovskite cathode for highly efficient CO2 electrolysis. J. Mater. Chem. 2019, 7, 22939–22949. [Google Scholar] [CrossRef] [Green Version]
  60. Xie, Z.; Zhao, H.; Chen, T.; Zhou, X.; Du, Z. Synthesis and electrical properties of Al-doped Sr2MgMoO6−δ as an anode material for solid oxide fuel cells. Int. J. Hydrogen Energy 2011, 36, 7257–7264. [Google Scholar] [CrossRef]
  61. Hou, M.; Sun, W.; Li, P.; Feng, J.; Yang, G.; Qiao, J.; Wang, Z.; Rooney, D.; Feng, J.; Sun, K. Investigation into the effect of molybdenum-site substitution on the performance of Sr2Fe1.5Mo0.5O6−δ for intermediate temperature solid oxide fuel cells. J. Power Sources 2014, 272, 759–765. [Google Scholar] [CrossRef]
  62. Wu, Y.; Wang, S.; Gao, Y.; Yu, X.; Jiang, H.; Wei, B.; Lü, Z. In situ growth of copper-iron bimetallic nanoparticles in A-site deficient Sr2Fe1.5Mo0.5O6−δ as an active anode material for solid oxide fuel cells. J. Alloys Compd. 2022, 926, 166852. [Google Scholar] [CrossRef]
  63. Cai, H.; Zhang, L.; Xu, J.; Huang, J.; Wei, X.; Wang, L.; Song, Z.; Long, W. Cobalt–free La0.5Sr0.5Fe0.9Mo0.1O3−δ electrode for symmetrical SOFC running on H2 and CO fuels. Electrochim. Acta 2019, 320, 134642. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the as-synthesized (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ samples.
Figure 1. XRD patterns of the as-synthesized (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ samples.
Materials 15 08268 g001aMaterials 15 08268 g001b
Figure 2. XRD patterns recorded for reduced (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ oxides in 5 vol.% H2/Argon at 1200 °C for 10h.
Figure 2. XRD patterns recorded for reduced (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ oxides in 5 vol.% H2/Argon at 1200 °C for 10h.
Materials 15 08268 g002
Figure 3. SEM micrographs of reduced (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ powder.
Figure 3. SEM micrographs of reduced (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ powder.
Materials 15 08268 g003aMaterials 15 08268 g003b
Figure 4. In-situ XRD measurements of oxidizing reduced (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ materials.
Figure 4. In-situ XRD measurements of oxidizing reduced (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ materials.
Materials 15 08268 g004aMaterials 15 08268 g004b
Figure 5. XRD patterns recorded at 30 °C and 850 °C in air for reduced (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ.
Figure 5. XRD patterns recorded at 30 °C and 850 °C in air for reduced (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ.
Materials 15 08268 g005
Figure 6. Thermal expansion behaviour of (a) oxidizing reduced Sr2Fe1.4TixMo0.6−xO6−δ sinters in air; (b) oxidized Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) samples in air; (c) reducing oxidized Sr2Fe1.4Ti0.1Mo0.5O6−δ and the reduced sample in 5 vol. % H2/argon.
Figure 6. Thermal expansion behaviour of (a) oxidizing reduced Sr2Fe1.4TixMo0.6−xO6−δ sinters in air; (b) oxidized Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) samples in air; (c) reducing oxidized Sr2Fe1.4Ti0.1Mo0.5O6−δ and the reduced sample in 5 vol. % H2/argon.
Materials 15 08268 g006
Figure 7. Oxygen content change of Sr2Fe1.4TixMo0.6−xO6−δ materials (a) in air and (b) in 5 vol.% H2 in argon; (c) results of DSC measurements for Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides in 5 vol.% H2/Ar.
Figure 7. Oxygen content change of Sr2Fe1.4TixMo0.6−xO6−δ materials (a) in air and (b) in 5 vol.% H2 in argon; (c) results of DSC measurements for Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides in 5 vol.% H2/Ar.
Materials 15 08268 g007
Figure 8. (a) Electrical conductivity of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides in air and 5 vol.% H2/Ar; (b) transport coefficients D and k determined as a function of temperature for Sr2Fe1.4Ti0.1Mo0.5O6−δ sample; (c) an exemplary normalized relaxation profile with fitting for Sr2Fe1.4Ti0.1Mo0.5O6−δ sinter.
Figure 8. (a) Electrical conductivity of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides in air and 5 vol.% H2/Ar; (b) transport coefficients D and k determined as a function of temperature for Sr2Fe1.4Ti0.1Mo0.5O6−δ sample; (c) an exemplary normalized relaxation profile with fitting for Sr2Fe1.4Ti0.1Mo0.5O6−δ sinter.
Materials 15 08268 g008aMaterials 15 08268 g008b
Figure 9. XRD patterns recorded for (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ with Ce0.8Gd0.2O1.9 electrolyte, after firing in air at 1200 °C for 10 h.
Figure 9. XRD patterns recorded for (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ with Ce0.8Gd0.2O1.9 electrolyte, after firing in air at 1200 °C for 10 h.
Materials 15 08268 g009
Figure 10. XRD patterns recorded for (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ with La0.8Sr0.2Ga0.8Mg0.2O3−δ electrolyte, after sintering at 1200 °C for 10 h in air.
Figure 10. XRD patterns recorded for (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ; (b) Sr2Fe1.4Ti0.2Mo0.4O6−δ and (c) Sr2Fe1.4Ti0.3Mo0.3O6−δ with La0.8Sr0.2Ga0.8Mg0.2O3−δ electrolyte, after sintering at 1200 °C for 10 h in air.
Materials 15 08268 g010
Figure 11. XRD patterns recorded for (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ with Ce0.8Gd0.2O1.9; (b) Sr2Fe1.4Ti0.1Mo0.5O6−δ with La0.8Sr0.2Ga0.8Mg0.2O3−δ electrolyte, after firing for 100h at 800 °C in air.
Figure 11. XRD patterns recorded for (a) Sr2Fe1.4Ti0.1Mo0.5O6−δ with Ce0.8Gd0.2O1.9; (b) Sr2Fe1.4Ti0.1Mo0.5O6−δ with La0.8Sr0.2Ga0.8Mg0.2O3−δ electrolyte, after firing for 100h at 800 °C in air.
Materials 15 08268 g011
Table 1. Rietveld refinement results including unit cell parameters for the as-synthesized and reduced Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides.
Table 1. Rietveld refinement results including unit cell parameters for the as-synthesized and reduced Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides.
Compositionx = 0.1x = 0.2x = 0.3
* As SynthesizedReducedAs SynthesizedReducedAs SynthesizedReduced
space groupPm-3mPm-3mPm-3mPm-3mPm-3mPm-3m
a [Å]3.9190 (1)3.9257 (1)3.9121 (1)3.9277 (1)3.9038 (1)3.9186 (1)
V3]60.19 (1)60.50 (1)59.88 (1)60.59 (1)59.49 (1)60.17 (1)
relative volume change ∆V0.52%1.19%1.14%
density [g/cm3]5.555.525.515.455.485.42
CHI23.253.532.322.772.044.64
Rp (%)1.381.651.261.461.151.53
Rwp (%)1.992.391.712.141.612.51
* With around 3.6% SrMoO4 phase.
Table 2. Thermal expansion coefficients TEC [10−6 K−1] of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) sinters from dilatometry studies and HT-XRD studies in air.
Table 2. Thermal expansion coefficients TEC [10−6 K−1] of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) sinters from dilatometry studies and HT-XRD studies in air.
HT-XRD
(400–850 °C)
Dilatometer
(400–900 °C, Oxidation in Air)
Dilatometer (400–900 °C, in Air for Oxidized Sinters)Dilatometer (30–900 °C in 5 vol. % H2/Argon for Reduced Sinters)
Sr2Fe1.4Ti0.1Mo0.5O6−δ17.416.415.3-
Sr2Fe1.4Ti0.2Mo0.4O6−δ19.516.719.513.7
Sr2Fe1.4Ti0.3Mo0.3O6−δ22.120.022.1-
Table 4. Structural parameters of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides with Ce0.8Gd0.2O1.9 and with La0.8Sr0.2Ga0.8Mg0.2O3−d electrolytes from 50:50 wt.% mixtures fired in air for 10h at 1200 °C.
Table 4. Structural parameters of Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) oxides with Ce0.8Gd0.2O1.9 and with La0.8Sr0.2Ga0.8Mg0.2O3−d electrolytes from 50:50 wt.% mixtures fired in air for 10h at 1200 °C.
Compositionx = 0.1Ce0.8Gd0.2O1.9x = 0.2Ce0.8Gd0.2O1.9x = 0.3Ce0.8Gd0.2O1.9
space groupPm-3mFm-3mPm-3mFm-3mPm-3mFm-3m
a [Å]3.9250 (1)5.4263 (1)3.9134 (1)5.4259 (1)3.9049 (1)5.4256 (1)
V3]60.47 (1)159.78 (1)59.93 (1)159.74 (1)59.54 (1)159.72 (1)
CHI23.294.353.08
Rp (%)1.851.951.76
Rwp (%)2.703.092.62
Compositionx = 0.1La0.8Sr0.2Ga0.8Mg0.2O3−δx = 0.2La0.8Sr0.2Ga0.8Mg0.2O3−δx = 0.3La0.8Sr0.2Ga0.8Mg0.2O3−δ
space groupPm-3mPm-3mPm-3mPm-3mPm-3mPm-3m
a [Å]3.9214 (1)3.9140 (1)3.9132 (1)3.9132 (1)3.9039 (1)3.9142 (1)
V3]60.30 (1)59.96 (1)59.92 (1)59.92 (1)59.50 (1)59.97 (1)
CHI22.031.982.23
Rp (%)1.661.661.70
Rwp (%)2.222.242.39
Table 5. Structural parameters of Sr2Fe1.4Ti0.1Mo0.5O6−δ sample with Ce0.8Gd0.2O1.9 and with La0.8Sr0.2Ga0.8Mg0.2O3−d electrolytes after the long-term studies for 100 h at 800 °C in air.
Table 5. Structural parameters of Sr2Fe1.4Ti0.1Mo0.5O6−δ sample with Ce0.8Gd0.2O1.9 and with La0.8Sr0.2Ga0.8Mg0.2O3−d electrolytes after the long-term studies for 100 h at 800 °C in air.
CompositionSr2Fe1.4Ti0.1Mo0.5O6−δCe0.8Gd0.2O1.9Sr2Fe1.4Ti0.1Mo0.5O6−δLa0.8Sr0.2Ga0.8Mg0.2O3−d
space groupPm-3mFm-3mPm-3mPm-3m
a [Å]3.9268 (1)5.4289 (1)3.9229 (1)3.9148 (1)
V3]60.55 (1)160.00 (1)60.37 (1)60.00 (1)
CHI22.513.62
Rp (%)2.021.93
Rwp (%)2.892.82
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zheng, K.; Albrycht, M.; Chen, M.; Qi, K.; Czaja, P. Tailoring the Stability of Ti-Doped Sr2Fe1.4TixMo0.6−xO6−δ Electrode Materials for Solid Oxide Fuel Cells. Materials 2022, 15, 8268. https://doi.org/10.3390/ma15228268

AMA Style

Zheng K, Albrycht M, Chen M, Qi K, Czaja P. Tailoring the Stability of Ti-Doped Sr2Fe1.4TixMo0.6−xO6−δ Electrode Materials for Solid Oxide Fuel Cells. Materials. 2022; 15(22):8268. https://doi.org/10.3390/ma15228268

Chicago/Turabian Style

Zheng, Kun, Maciej Albrycht, Min Chen, Kezhen Qi, and Paweł Czaja. 2022. "Tailoring the Stability of Ti-Doped Sr2Fe1.4TixMo0.6−xO6−δ Electrode Materials for Solid Oxide Fuel Cells" Materials 15, no. 22: 8268. https://doi.org/10.3390/ma15228268

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