Yao
Lu
a,
Hailei
Zhao
*ab,
Kui
Li
a,
Xuefei
Du
a,
Yanhui
Ma
a,
Xiwang
Chang
c,
Ning
Chen
a,
Kun
Zheng
de and
Konrad
Świerczek
de
aSchool of Materials Science and Engineering, University of Science and Technology Beijing, Beijing 100083, China. E-mail: hlzhao@ustb.edu.cn; Fax: +86 10 82376837; Tel: +86 10 82376837
bBeijing Municipal Key Laboratory of New Energy Materials and Technologies, Beijing 100083, China
cState Key Laboratory of Advanced Metallurgy, University of Science and Technology Beijing, Beijing 100083, China
dDepartment of Hydrogen Energy, Faculty of Energy and Fuels, AGH University of Science and Technology, al. A. Mickiewicza 30, 30-059 Krakow, Poland
eAGH Centre of Energy, AGH University of Science and Technology, ul. Czarnowiejska 36, 30-054 Krakow, Poland
First published on 28th March 2017
A cost-effective doping strategy was developed to enhance the oxygen permeability and structural stability of BaFeO3−δ. We demonstrated that the alkaline earth metal element Ca, which is usually considered an A-site dopant for perovskite oxides, can be successfully introduced into the B-site of BaFeO3−δ. The cubic perovskite structure of BaFe1−xCaxO3−δ was stabilized down to room temperature for the Ca-doping concentration range from 5 to 15 at%. First principles calculations not only proved the preference of Ca at the B-site with lower defect formation energies than the A-site, but also demonstrated that the migration of the oxygens located greater distances from the Ca position is characterized by lower barrier energies than those in the Ca vicinity and even lower than that for the undoped BaFeO3−δ. We found that these favourable, low energy barrier paths away from the Ca sites exert more pronounced effects on the oxygen migration at diluted dopant concentrations, and hence, the material with x = 0.05 level of substitution shows a higher oxygen permeability with a lower activation energy compared to the undoped or highly-doped BaFeO3−δ. The BaFe0.95Ca0.05O3−δ membrane is characterized by a high oxygen permeability of 1.30 mL cm−2 min−1 at 950 °C and good long-term stability at 800/900 °C, as obtained over 200 h. Therefore, the feasibility and applicability of Ca-doping at the B-site of the perovskite can be highlighted, which allows for the enhancement of the oxygen migration ability, originating from the appropriate tuning of the lattice structure.
A dense ceramic oxygen-permeable membrane was firstly reported by Teraoka et al. on a MIEC-type perovskite oxide with La1−xSrxCo1−yFeyO3−δ composition in the 1980s.9 With an increasing oxygen vacancy concentration and a decreasing average bond strength upon introduction of Sr and Co at A- and B-sites, respectively, the oxygen permeability was found to increase and reach the highest flux value of 3.1 mL cm−2 min−1 for SrCo0.8Fe0.2O3−δ material at 850 °C. These results have attracted considerable attention to cobalt-based perovskite-type oxygen ionic transport membranes during the past decades. Recent investigations on mixed-conducting oxide materials and membranes with various A/B-site compositions have shown that Ba0.5Sr0.5Co0.8Fe0.2O3−δ and BaCo0.7Fe0.2Nb0.1O3−δ oxides are the most promising candidates, on an account of their high oxygen permeability, which takes advantage of the fact that the A-site is occupied by large Ba2+ cations.10,11 However, these cobalt-based materials always suffer from a structural instability and a corresponding oxygen permeability degradation during long term operation, which have been ascribed as originating from the reduction of cobalt ions under the low oxygen partial pressure atmospheres at elevated temperatures.12–14 Furthermore, the high price of cobalt is another factor that may limit the widespread application of such cobalt-based membranes. Therefore, development of cobalt-free materials for MIEC membrane applications seems very important.
Already a large number of cobalt-free membranes have been developed to improve the reduction resistance, such as BaFe1−xCexO3−δ, Ba0.5Sr0.5Fe0.8Zn0.2O3−δ, La1−xSrxCr1−yFeyO3−δ.15–17 Considering the oxygen permeability in conjunction with the stability in the reducing atmosphere, BaFeO3−δ is believed to be one of the most promising cobalt-free MIECs. The merits of BaFeO3−δ are strongly associated with the presence of Fe at the B-site, which has a higher B–O bond strength, as well as a full occupation of the A-site by Ba cations, allowing it to sufficiently expand the lattice structure, which is beneficial for oxygen migration.18–22 However, since the Shannon radii-based tolerance factor (t) is higher than 1, which is a result of a large mismatch between the size of the A- and B-site cations, BaFeO3−δ has been found to maintain the oxygen vacancy-disordered, cubic perovskite structure only at temperatures higher than 825 °C. At lower temperatures, a phase transformation from cubic to a low symmetry structure occurs and is accompanied by a large volume change, as well as a seriously declined oxygen permeability.18,22
In order to stabilize the cubic structure of BaFeO3−δ down to lower temperatures, various dopants have been introduced to either the A-site or B-site, such as at the A-site: La, Ca, Sm and at the B-site: Ce, La, Y, Cu, Ni, Zr, Ta, In, Gd, Nb.18–29 For tuning the lattice structure to cubic, A-site dopants are always chosen to be smaller than Ba cations, while B-site dopants are selected to be larger than Fe cations, according to the required decrease of the tolerance factor (t).30 For the two doping strategies, the B-site doping with larger cations is more preferred because the expanded lattice is beneficial for oxygen ion diffusion, which is in contrast to the A-site doping with smaller cations than Ba.20,22,28 Most of the promising candidates for the B-site dopants are rare earth elements because of their common moderate valence (+3), relatively large ionic radius, and strong reduction resistance in comparison to the transition metal elements, which lead to a high oxygen vacancy concentration, stable perovskite cubic structure and good structural stability, respectively.22,28 Despite the advantages of the rare earth dopants, the higher bond strength between such metals and oxygen anions, compared to that of the Fe–O bond, is unfavourable for oxygen migration in the lattice. It also should be mentioned that the high price and scarce resources of rare earth metals may seriously restrict applications of such MIECs in industrial, large-scale applications.
Herein, we report for the first time a novel BaFeO3−δ-based perovskite, doped with the cheap and abundant alkaline earth metal Ca2+ at the B-site. Usually, Ca2+ is regarded as an A-site dopant because of its similar properties with respect to Ba, being in the same main group of the periodic table. However, considering the much smaller ionic radius of Ca2+, similar in the BO6 octahedron to the rare earth elements, it appears to be a very promising potential B-site dopant for tuning the lattice structure of BaFeO3−δ. The advantages of Ca2+ include its low valence state (+2) and a relatively weak Ca–O bond strength compared with the rare earth element–oxygen bonds, and it would endow the host BaFeO3−δ material with a high oxygen vacancy concentration and a low oxygen migration energy. As a result, a high oxygen permeation flux through such designed membranes can be expected. In this article, the effects of Ca-doping on the lattice structure, oxygen nonstoichiometry, electrical conductivity, oxygen surface exchange and bulk diffusion, oxygen permeability and long term stability were systematically investigated. First principles calculations were also conducted to explore the location preference of Ca atoms and to understand the oxygen migration behaviour in the materials. The experimental results confirm the anticipated superior oxygen permeability of doped BaFe1−xCaxO3−δ for a low-level doping degree, which stems from a complex influence of Ca on the crystal structure, oxygen nonstoichiometry and transport properties. Abundant resources and the low price of all the elements as well as the recorded excellent performance of BaFe1−xCaxO3−δ make the elaborated membrane materials very competitive for practical industry applications.
The microstructure and element distribution on the surface of the sintered membranes (x = 0.05 and 0.15) were analysed by scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDXS) using a LEO-1450 scanning electron microscope equipped with a spectrometer (Noran System 7, Thermo Scientific, USA) at excitation voltages of 20 kV for SEM and EDXS analyses. The electron probe microanalysis (EPMA, JEOL, JXA-8100) was also conducted to analyse the chemical composition of the sintered membranes (x = 0.05 and 0.15). X-ray photoelectron spectroscopy (XPS) data were obtained using a Thermo ESCALAB 250X (USA) electron spectrometer with a 150 W Al Kα X-ray source.
The temperature-dependent weight of the BaFe1−xCaxO3−δ (x = 0.025, 0.05 and 0.15) samples was characterized by thermogravimetric analysis (TG) on a TA Instruments Q5000 IR (USA) apparatus in air or argon with a flow rate of 100 mL min−1 from 25 to 900 °C. According to the TG results and the initial oxygen nonstoichiometry, δ0, which was evaluated with iodometric titration method,34 the oxygen nonstoichiometry δt at high temperatures was calculated using eqn (1),
![]() | (1) |
The electrical conductivity of the considered materials was measured by a four-terminal DC method in air from 200 °C to 900 °C. The chemical bulk diffusion coefficient (Dchem) and surface exchange coefficient (kchem) were determined with the electrical conductivity relaxation (ECR) method. Before measurements, a dense sinter with bar shape was placed in a furnace and tested at temperatures between 650 °C and 850 °C with an interval of 50 °C. The material was supplied with a 10 vol% O2/N2 gas mixture at a constant flow of 200 mL min−1 for about 1 h to reach the steady-state. By abruptly switching the atmosphere from 10 vol% to 20 vol% O2/N2 gas mixture, the electrical conductivity changed due to the variation of the oxygen partial pressure, and the dependence on time was recorded by a four-terminal DC method using a high-precision digital multimeter (Keithley 2100). Then, the oxygen ion diffusion and surface exchange coefficients were obtained by fitting the recorded electrical conductivity relaxation curves based on Fick's second law.35–37
![]() | (2) |
![]() | ||
Fig. 1 XRD patterns of BaFe1−xCaxO3−δ samples (x = 0.025–0.15) after sintering in air at 1200 °C for 4 h. |
For further exploration of the location preference of Ca in the cubic perovskite BaFeO3−δ oxide, first principles calculations were employed to identify the energy cost of substitutional defect formation for doping of Ca at A- or B-sites. The defect formation energies (Eform) were calculated using the following eqn (3),
Eform = Edefect − Ehost + μhost − μdope | (3) |
Ba8Fe8O23-1 | Ba8Fe8O23-2 | Ba8Fe8O23-3 | |
---|---|---|---|
E(CaBa)/eV | 1.19 | 1.00 | — |
E(CaFe)/eV | 0.81 | 0.10 | 0.41 |
Besides, to further confirm the distribution of Ca in the samples, distribution of the elements was studied by EPMA and EDXS methods for compounds with x = 0.05 and 0.15. Both results (Tables S2 and S3†) indicate that the combined molar fraction of Fe and Ca is similar to that of Ba. With an increase of the Ca-doping content to 0.15, the molar fraction of Ca increases to three times that of the x = 0.05 sample, which follows the designed cation ratios in the materials. Moreover, the symmetric nature of the Ca 2p XPS spectrum, as shown in Fig. S3d,† demonstrates only one kind of chemical environment of Ca, which is most likely in the perovskite phase. In addition, the membranes show dense microstructures (Fig. 2a and e), and the EDXS mapping analysis (Fig. 2) reveals a homogeneous distribution of Ba, Ca and Fe elements, with no phase segregation detected up to the resolution of the method. Considering that some secondary phase should emerge if the Ca dopant would occupy the A-site rather than the B-site, the above results with a combination of structural chemical analysis and first principles calculations, as well as additional studies (e.g. attempt to synthesize Ba0.85Ca0.15FeO3−δ, Fig. S4†), provide strong support to Ca-doping at the B-site of the BaFe1−xCaxO3−δ lattice.
In order to assess the phase stability of BaFe1−xCaxO3−δ materials at high temperatures, in situ XRD measurements were conducted on powder samples with x = 0.025, 0.05 and 0.15 under an air atmosphere in a cooling process within the temperature range from 900 °C to room temperature. The representative results are presented with the corresponding PDF standard cards of the related structures in Fig. 3. A phase transition for BaFe0.975Ca0.025O3−δ material from high-temperature cubic to triclinic perovskite symmetry was found to occur below 600 °C, as shown in Fig. 3a. A single phase cubic perovskite phase for samples with x = 0.05 and 0.15 (Fig. 3b and S5†) is maintained throughout the whole temperature range, and no phase transition could be evidenced. The XRD patterns of both materials were refined using the Rietveld method by assuming the presence of Ca at the A-site or B-site, respectively. Better results (smallest residues) were obtained if Ca was placed at the Fe site (Table S4†). These results also support previous findings that the Ca dopant is substituting the Fe, rather than the barium site.
![]() | ||
Fig. 3 In situ XRD patterns of BaFe1−xCaxO3−δ samples with x = 0.025 (a) and x = 0.05 (b) collected from high to room temperature. |
The lattice parameters of the two cubic samples (x = 0.05 and 0.15) were calculated based on the Rietveld refinement of the in situ XRD results (Fig. 3b and S5†), and the results are plotted in Fig. S6.† The values increase with temperature, and the dependence appears to be accelerated when the temperature is higher than 400 °C. This is associated with the so called chemical expansion known to appear for such MIEC-type materials. The effect adds to the basic thermal expansion and originates from the release of the lattice oxygen, causing repulsion of two of the B-site cations in the immediate vicinity of a generated oxygen vacancy, as well as an associated increase in the Fe radius with reduced valence.52 By applying the linear fitting to the temperature-dependent lattice parameter data in low (25–300 °C) and high (600–900 °C) temperature ranges, the linear expansion coefficients (TEC) could be obtained as 17.8 × 10−6 and 29.0 × 10−6 K−1 for x = 0.05 sample, and as 14.9 × 10−6 and 38.8 × 10−6 K−1 for x = 0.15 material (shown in inset of Fig. S6†), respectively. Such results are comparable with other reported BaFeO3−δ-based materials.19,53 Apparently, Ca-doping decreases the TECs in the low temperature range, while it exerts an opposite effect at the high temperature range. The decreased TEC at low temperatures is most likely due to a decrease in the oxygen deficiency with Ca-doping, as is presented in Fig. 4. Such influence is expected to decrease the anharmonicity of atomic vibrations.54 On the other hand, the increased TECs in the high temperature range mostly result from larger changes of the oxygen deficiency with Ca-doping between 600 and 900 °C, as evidenced in the TG results (Fig. 4). The impact of Ca-doping on the thermal reducibility of BaFe1−xCaxO3−δ was further analysed by TG (Fig. 4a). It should be mentioned that since there are some adsorbates being adsorbed on the material's surface, all samples exhibited a weight loss below 300 °C at the first heating process (Fig. S7†).27 During the following cycles, this kind of extrinsic weight change was eliminated, and thus, data from the second thermal cycle could be taken into consideration for oxygen nonstoichiometry calculations (Fig. 4b). The weight of the samples with x = 0.05 and 0.15 starts to decrease from ∼300 °C, which is ascribed to the thermal reduction of Fe ions and the corresponding release of lattice oxygen. This is in agreement with the previously mentioned variation of the temperature-dependent lattice parameter plots (Fig. S6†). It is worth to mention that there is an unusual weight loss and recovery in the middle temperature range for the sample with x = 0.025, which might be related to the structural transition (Fig. 3a) or possibly to some other transformation related to the electric or magnetic properties of the material. While this phenomenon was also observed in repeated studies, further tests are needed to elucidate its nature.
Combining the TG results with the room temperature oxygen nonstoichiometry obtained by an iodometric titration technique, the oxygen vacancy concentrations at higher temperatures were calculated and plotted in Fig. 4b. The values were found to decrease with the increasing Ca-doping content (Table S5†), while the calculated average valence of Fe increases, suggesting a charge compensation mechanism by varying the B-site element valence, rather than generating more oxygen vacancies. The defects produced by the acceptor-type Ca-doping in BaFeO3−δ can be expressed by eqn (4)–(6):
![]() | (4) |
![]() | (5) |
![]() | (6) |
Such acceptor doping induces the generation of electronic holes for charge compensation (eqn (4)), while the stronger Ca–O bond, compared to the Fe–O bond, causes the decrease of the concentration of the oxygen vacancies, but a concomitant increase in the amount of the electronic holes (eqn (5)). Those electronic holes can be expressed as the increase of the valence of Fe cations (eqn (6)). Such a charge compensation mechanism might be associated with the reduced lattice distortion by substitution of larger Ca and doping for Fe cations, which makes Fe ions not necessarily increase their radius by reduction of the valence, and therefore, for mitigating lattice distortion.
The Ca-doping also has impacts on the electronic transport performance of BaFe1−xCaxO3−δ. The temperature dependence of the electrical conductivity of BaFe1−xCaxO3−δ in the Arrhenius plot is shown in Fig. 5. For most of the MIECs, the electrical conductivity mostly comes from the electronic component, since its magnitude is usually at least one or two orders larger than that of ionic conductivity.23 In the corresponding Arrhenius coordinates, the electronic conductivity increases linearly with the increase of the temperature in a lower temperature range, indicating a Zener double-exchange mechanism via B–O–B bonds. At higher temperatures, the electronic conductivity deviates from the former linear trend, showing a fast decrease. This is associated with the lattice oxygen release at those temperatures, which would cause a partial annihilation of the electrical holes, and thus, decrease the concentration of the p-type charge carriers, as described by eqn (7).
![]() | (7) |
![]() | ||
Fig. 5 Temperature dependence of electrical conductivity of BaFe1−xCaxO3−δ (x = 0.025, 0.05, 0.10 and 0.015) samples in air, represented as an Arrhenius-type plot. |
In addition, increasing the oxygen vacancy concentration breaks the B–O–B charge transfer, as the oxygen is needed for the double exchange mechanism to occur.52 Furthermore, Fig. 5 also reveals that the substitution of redox-inactive Ca cations for Fe at the B-site decreases the electronic conductivity, even though the Ca substitution leads to a higher average valence of Fe (Table S5†). This can be understood considering that Ca-doping also severs the B–O–B conduction path, with the inability of Ca 3d orbitals to connect with the O 2p states, which leads to breaking the conduction network.
The oxygen bulk diffusion coefficients (Dchem) and surface exchange coefficients (kchem) of the BaFe1−xCaxO3−δ materials with x = 0.025–0.15 were measured by ECR method in the temperature range of 650–850 °C. Fig. 6a shows the typical ECR curves for the BaFe1−xCaxO3−δ dense samples being collected after a sudden change of the oxygen partial pressure from 0.1 to 0.2 atm at 800 °C. Based on ECR theory,35–37 the oxygen bulk diffusion coefficients (Dchem) and surface exchange coefficients (kchem) of the tested materials were calculated by fitting the ECR data. Both of the coefficients were found to decrease with the Ca-doping content, as shown in Fig. 6b and c. The calculated activation energies, extracted from the corresponding Arrhenius plots of the temperature dependence of Dchem and kchem, exhibit the lowest value for the x = 0.025 sample, and increase with the Ca-doping content, implying that the oxygen bulk migration and surface oxygen exchange process are hindered by introducing more Ca cations into the B-sites. The lower activation energy of kchem than Dchem is probable due to the high activity of the oxygen reduction reaction of BaFeO3−δ based materials,55 which is also found in some other similar perovskite-type MIEC materials.56 Nevertheless, it is worth mentioning that the x = 0.025 sample exhibits a very fast oxygen bulk diffusion and surface exchange kinetics, with the coefficients of 1.21 × 10−4 cm2 s−1 and 3.97 × 10−3 cm s−1 at 700 °C, respectively, which is among the highest level of many advanced perovskite MIECs.36,53
In order to advance the understanding of the Ca-doping effect on oxygen migration behaviour, first principles calculations were applied to elucidate the oxygen vacancy transport energy through different routes in the crystal lattice, in which a Ca atom was incorporated into a 2 × 2 × 2 BaFeO3 supercell by substituting one Fe atom. The BaFeO3 pristine cell was also constructed for comparison (Fig. 7a). It shows only one equivalent oxygen vacancy migration path in the lattice, of which the corresponding oxygen migration energy is 0.91 eV.28 In the Ca-doped lattice, the coordination environment of the oxygen vacancy is obviously diversified. There are three possible sites for oxygen vacancies: nearest, second nearest and third nearest neighbours to the Ca dopant. They are donated as VO1, VO2 and VO3, respectively, and depicted in Fig. 7b, which are equivalent to the CaFe models of Ba8Fe7CaO23-1, Ba8Fe7CaO23-2, Ba8Fe7CaO23-3 in Fig. S2.† Those models, at first, were separately optimized in terms of geometry, which was successfully done with the energy reaching the respective minimums. It was found that the state with the oxygen vacancy at the VO1 position possesses the highest potential energy compared to the other two states, as shown in Fig. 7c, implying an unstable state for the location in the vicinity of the Ca2+ cations. The other two sites are more favourable for the presence of the oxygen vacancies. This finding could also be rationalized by a measured decrease of the oxygen vacancy concentration with the increasing Ca-doping content, as previously outlined in Fig. 4b.
From a geometry point of view, there are five primitive oxygen migration paths between the three kinds of oxygen vacancies and neighbouring oxygen atoms, as shown in Fig. 7c. The individual oxygen migration energy for different paths is calculated and listed in Table 2. Among the five paths, the largest value of the migration barrier energy (1.82 eV) occurs when the oxygen anions migrate along the path closest to the Ca cations, which might be correlated with a high basicity of calcium and may limit the removal of oxygen anions with high acidity. The increase in the activation energy of Dchem, as mentioned in Fig. 6b, can be attributed to the increased proportion of high barrier energy paths with a higher Ca-doping content. Nevertheless, the barrier energies along other migration paths exhibit much lower values than that in pristine BaFeO3−δ (0.91 eV). Since the lattice structure is expanded by doping the larger Ca cations, the barrier energies of the other migration paths, away from Ca, could probably be decreased by crossing the bottleneck of Ba–Fe–Ba in the cubic perovskite lattice. Thus, considering the coexistence of both promoted and impeded oxygen migration paths, there should be an optimal Ca-doping content, at which the cubic structure can be maintained under a minimized restriction on the oxygen ion migration.
Migration barrier energy [eV] | |||
---|---|---|---|
VO1 | VO2 | VO3 | |
VO1 | 1.82 | 0.89 | |
VO2 | 0.19 | 0.49 | 0.36 |
VO3 | 0.66 | 0.50 |
The oxygen permeation fluxes of the BaFe1−xCaxO3−δ membranes (x = 0–0.15) with a 1.0 mm thickness were measured at temperatures between 800 and 950 °C using pure He as the sweep gas (Fig. 8a). For all doped membranes, a linearly increasing trend of the oxygen permeability with temperature was observed, indicating a thermal activation process of the oxygen permeation. At the same time, the oxygen permeation flux of the pristine BaFeO3−δ membrane exhibits a very low value at low temperatures and an abrupt increase around ∼825 °C. This is ascribed to the phase transition of BaFeO3−δ from the oxygen vacancy ordered hexagonal structure to the disordered cubic perovskite structure.34 Because as evidenced in Fig. 3a, the phase transition temperature of the x = 0.025 sample is decreased by B-site substitution to ∼600 °C, and in the range of 800–950 °C, the membrane shows a gradual increase in the oxygen permeation flux. However, this membrane with x = 0.025 doping content delivers the highest oxygen permeability of 1.39 mL cm−2 min−1 at 950 °C, which is considerably high for the cobalt-free membrane. Due to the increased activation energy of the oxygen bulk diffusion (Fig. 6b) and the reduced oxygen vacancy concentration (Table S5†), a further increase in the Ca-doping level leads to a gradual decrease in the oxygen permeability.
The activation energies of the oxygen permeability of the x = 0–0.15 samples were also calculated from the corresponding Arrhenius-type temperature dependent oxygen permeation plots according to the Wagner equation.22,25 The obtained dependence of the activation energies and the oxygen permeability at 900 °C on Ca-doping content is depicted in Fig. 8b. In contrast to the changes of the oxygen permeability, the opposite behaviour can be seen for the activation energy of the oxygen permeation, which markedly decreases first and then increases gradually with the Ca-doping content. Compared with pure BaFeO3−δ, the lower activation energy of the slightly doped samples x = 0.025 and 0.05 can be ascribed to the expanded lattice structure, caused by the large-size Ca substitution and the resulting favourable transport paths away from the Ca position, as discussed in the section for the first principles calculations (Fig. 7). However, with a further increase in the doping content, the growing proportion of the high-barrier paths around the Ca positions outweighs the positive effect of other low energy barrier paths and leads to the increase of the activation energy. Thus, slight doping (x = 0.025, 0.05) is beneficial to maximize the advantages of the low barrier paths, and in consideration for the effective stabilization of the cubic perovskite structure in the whole temperature range of 25 to 950 °C, the x = 0.05 dopant level is preferred as the most suitable one. It should be mentioned that the apparent activation energy of the oxygen permeation for the x = 0.05 membrane (0.34 eV) is lower than that of the Dchem (0.92 eV). The inconsistency is likely coming from the different testing methods with non-identical conditions.57,58 This effect was also shown in other works.53,59,60
The long-term stability of the x = 0.05 membrane with 0.6 mm thickness was then examined at 800/900 °C under a He/Air gradient. The oxygen permeability is stable for 200 h with no obvious deterioration, as shown in Fig. 9, indicating the excellent structural stability of the BaFe0.95Ca0.05O3−δ material. The tested x = 0.05 membrane was also subjected to XRD examination to evaluate its surface structure change after long-term operation, and there were no impurity phases observed, apart from the original cubic perovskite phase, as shown in Fig. S8.† In contrast to the easy reduction and decomposition of cobalt-based perovskite membranes in a low partial pressure atmosphere,12–14 the Ca-doped cobalt-free BaFe0.95Ca0.05O3−δ membrane displays a high structural stability with good long term operational reliability. These results strongly suggest that Ca-doped BaFeO3−δ is a promising oxygen permeation membrane material, taking into account its excellent performance and cheap and abundant material resources.
![]() | ||
Fig. 9 Long term stability of BaFe0.95Ca0.05O3−δ membrane with 0.6 mm thickness for 100 h at either 900 or 800 °C. |
The resistance of oxygen permeable membranes against low oxygen partial pressure reduction and CO2 erosion is of great importance in practical industrial applications. The chemical and structural stability of the BaFe1−xCaxO3−δ membranes with x = 0.05 and 0.15 was evaluated by treating the fresh membranes in 5 vol% H2/Ar and in 2 vol% CO2/N2 at 800 °C for 8 h, respectively. The surface phase structure and morphology of the treated membranes were studied by XRD and SEM examinations. For samples treated in 5 vol% H2/Ar, even though the surfaces of both the membranes changed and some segregates are exsolved, as shown in Fig. S9a and b,† the surface impurities are undetectable, and the main phase structure is maintained (Fig. S10†), demonstrating that the surface damage is not so severe. It can be stated that the BaFe1−xCaxO3−δ materials have a good chemical and structural stability in reducing atmospheres. Furthermore, the SEM images indicate that Ca-doping mitigates the phase segregation and enhances the structural stability of materials. When treated in a CO2-containing atmosphere (2 vol% CO2/N2), some remarkable changes on the surface structure were observed. XRD analysis reveals precipitation of BaCO3 and BaFe2O4 impurity phases (Fig. S10†), while SEM observation shows an emergence of large segregates on the membrane surface (Fig. S9c and d†), indicating the weaker resistance towards CO2 erosion compared to H2 reduction. Nevertheless, the respective XRD peak intensities of the secondary phases decreased remarkably with the increasing Ca-doping level, suggesting that the B-site Ca-doping enhances the tolerance of BaFe1−xCaxO3−δ towards CO2.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c7ta00907k |
This journal is © The Royal Society of Chemistry 2017 |