Processing math: 100%
Shanshan Jiang, Hao Qiu, Shaohua Xu, Xiaomin Xu, Jingjing Jiang, Beibei Xiao, Paulo Sérgio Barros Julião, Chao Su, Daifen Chen, and Wei Zhou, Investigation and optimization of high-valent Ta-doped SrFeO3–δ as air electrode for intermediate-temperature solid oxide fuel cells, Int. J. Miner. Metall. Mater., 31(2024), No. 9, pp.2102-2109. https://dx.doi.org/10.1007/s12613-024-2872-1
Cite this article as: Shanshan Jiang, Hao Qiu, Shaohua Xu, Xiaomin Xu, Jingjing Jiang, Beibei Xiao, Paulo Sérgio Barros Julião, Chao Su, Daifen Chen, and Wei Zhou, Investigation and optimization of high-valent Ta-doped SrFeO3–δ as air electrode for intermediate-temperature solid oxide fuel cells, Int. J. Miner. Metall. Mater., 31(2024), No. 9, pp.2102-2109. https://dx.doi.org/10.1007/s12613-024-2872-1
Research Article

Investigation and optimization of high-valent Ta-doped SrFeO3–δ as air electrode for intermediate-temperature solid oxide fuel cells

Author Affilications
  • Corresponding author:

    Shanshan Jiang      E-mail: jss522@just.edu.cn

    Chao Su      E-mail: chao.su@just.edu.cn

  • To explore highly active and thermomechanical stable air electrodes for intermediate-temperature solid oxide fuel cells (IT-SOFCs), 10mol% Ta5+ doped in the B site of strontium ferrite perovskite oxide (SrTa0.1Fe0.9O3–δ, STF) is investigated and optimized. The effects of Ta5+ doping on structure, transition metal reduction, oxygen nonstoichiometry, thermal expansion, and electrical performance are evaluated systematically. Via 10mol% Ta5+ doping, the thermal expansion coefficient (TEC) decreased from 34.1 × 10–6 (SrFeO3–δ) to 14.6 × 10–6 K–1 (STF), which is near the TEC of electrolyte (13.3 × 10–6 K–1 for Sm0.2Ce0.8O1.9, SDC), indicates excellent thermomechanical compatibility. At 550–750°C, STF shows superior oxygen vacancy concentrations (0.262 to 0.331), which is critical in the oxygen-reduction reaction (ORR). Oxygen temperature-programmed desorption (O2-TPD) indicated the thermal reduction onset temperature of iron ion is around 420°C, which matched well with the inflection points on the thermos-gravimetric analysis and electrical conductivity curves. At 600°C, the STF electrode shows area-specific resistance (ASR) of 0.152 Ω·cm2 and peak power density (PPD) of 749 mW·cm–2. ORR activity of STF was further improved by introducing 30wt% Sm0.2Ce0.8O1.9 (SDC) powder, STF + SDC composite cathode achieving outstanding ASR value of 0.115 Ω·cm2 at 600°C, even comparable with benchmark cobalt-containing cathode, Ba0.5Sr0.5Co0.8Fe0.2O3–δ (BSCF). Distribution of relaxation time (DRT) analysis revealed that the oxygen surface exchange and bulk diffusion were improved by forming a composite cathode. At 650°C, STF + SDC composite cathode achieving an outstanding PPD of 1117 mW·cm–2. The excellent results suggest that STF and STF + SDC are promising air electrodes for IT-SOFCs.
  • Solid oxide fuel cells (SOFCs), due to the high energy conversion efficiency and good fuel flexibility, are treated as a promising next-generation energy technology [17]. The commercial utilization of SOFCs has been majorly restricted by their typical working temperature (>800°C), the high operating temperature causes many practical matters such as thermal & chemical compatibility between SOFC components, sealing issues, and tardiness start-up. However, the reduced operation temperature is often accompanied by severe polarization resistance loss and sluggish oxygen reduction reaction (ORR) activity. To realise the reduced temperature operation of SOFCs, for improved components durability and materials compatibility and accelerate the start-up process, nowadays there are tremendous research efforts focus on the air electrode materials with desirable ORR activity under lower temperature range [810].

    Cobalt-based oxides with mixed ionic and electronic conductivity (MIEC) are thoroughly investigated as air electrode materials for IT-SOFCs due to their expectative ORR activity and electrical conductivity. For example, Ba0.5Sr0.5Co0.8Fe0.2O3–δ (BSCF), SrCo0.8Nb0.1Ta0.1O3–δ (SCNT), and Sm0.5Sr0.5CoO3 (SSC) are epidemic MIEC air electrodes for IT-SOFCs. Attractively low values of electrode polarization resistance have been investigated for these cobalt-based cathodes at temperatures as low as 650°C [1113]. However, cobalt-based perovskites usually exhibited high thermal expansion coefficients (TECs) from the transition of Co3+ to Co4+, and the low spin to high spin. Thermal mismatch between the air electrode and electrolyte layer could lead to electrode stratification and reduce the working life of SOFCs [14].

    Recently, many efforts have been focused on overcoming this problem by developing high-performance cobalt-free perovskite oxide cathodes. As well known, SrFeO3–δ is a typical ABO3-type mixed-ionic electric conductor, it consists of both brownmillerite and cubic phases, thus restraining ORR kinetics and accompanied by incompatibility between cell components. A-site or B-site doped SrFeO3–δ based perovskite oxides, La1–xSrxFeO3–δ, Sr1–xBixFeO3–δ, SrNb0.1Fe0.9O3–δ (SNF), SrTi0.1Fe0.9O3–δ, and PrxSr1–xFeO3–δ have been investigated as potential cathode materials [1518]. However, those cobalt-free electrodes typically have lower TECs along with a trade-off electrochemical performance, and the TEC values are still much higher than electrolytes. Therefore, it is crucial to develop air electrode materials with enhanced ORR activity and prolong thermal stability for IT-SOFCs [1920]. In our previous work, high-valent metal cations Nb5+ was used as dopants to tailor the electroactivity of SrFeO3–δ, and demonstrated that adding a certain amount of Nb5+ to the B-site in SrFeO3–δ is an effective way to stabilize the cubic perovskite structure, thereby improving the ORR activity and stability. Because Nb5+ and Ta5+ are well known VB group transition metals, they have fixed +5 oxidation states and similar ionic radius, which is close to (0.64 Å). Therefore, we propose to doping Ta5+ into the B-site of SrFeO3–δ materials as a novel cathode material for IT-SOFC.

    Herein, we investigated 10mol% high-valent Ta doped SrFeO3–δ as potential air electrode material. SrTa0.1Fe0.9O3–δ (STF) shows extremely low TEC values and excellent polarization resistance. The ORR activity was further enhanced by introducing 30wt% SDC powder to form the SrTa0.1Fe0.9O3–δ + Sm0.2Ce0.8O1.9 (STF + SDC, 70:30 mass ratio) composite cathode. The area-specific resistance (ASR) of the STF + SDC composite cathode is 0.012, 0.019, 0.044, 0.115, and 0.233 Ω·cm2 at 750, 700, 650, 600, and 550°C, respectively. The ASR at 550°C is even slightly lower than the benchmark reduction electrode material, BSCF, which is 0.24 Ω·cm2 at 600°C under similar conditions.

    STF and SrFeO3–δ were synthesized via solid state method combined with high-energy ball milling. Stoichiometric amounts of SrCO3 (AR, Aladdin), Ta2O5 (AR, Aladdin), and Fe2O3 (AR, Aladdin) were ball milled, Pulverisette 6, FRITSCH (400 r/min, 30 min) in alcohol media. The precursor was further calcinated at 1200°C for 20 h twice to achieve a highly pure STF sample.

    The STF + SDC (mass ratio, 70:30) powder was premixed by high-energy ball-milling (400 r/min, 30 min). Symmetrical cells of STF |SDC| STF (or STF + SDC |SDC| STF + SDC) were fabricated by spraying STF (or STF + SDC) slurry symmetrically onto the SDC disks. The symmetrical cells with STF air electrodes were calcinated at 800, 900, and 1000°C in an ambient atmosphere for 2 h. Half cells with the configuration of SDC (~20 μm) | SDC–Ni (40:60 mass ratio) were constructed via dry pressing. The STF cathode was sprayed onto the SDC electrolyte layer and subsequently calcinated at 800°C for 2 h. The single fuel cell with STF + SDC composited cathode was fabricated using the same process.

    The X-ray diffraction (XRD) of STF, SF, and STF + SDC composite oxides were conducted by Bruker AXS D8 Advance. Field emission scanning electron microscope (FE-SEM, JEOL-S4800) was employed in detecting the microstructure of the electrode and the single cell. FEI Titan G2 F30 (300 kV) was utilized for high-resolution transmission electron microscopy (HR-TEM), high-angle annular dark field scanning TEM (HAADF-STEM), and energy dispersive X-ray spectroscopy (EDS) mapping. Oxygen temperature-programmed desorption (O2-TPD) was examined via self-building equipment, a programmed temperature furnace combined with a mass spectrometer. The room temperature oxygen nonstoichiometric (δ) in STF oxide was measured by iodometric titration. 100 mg STF powder was dissolved in HCl solution (0.6 mol·L–1) under nitrogen atmosphere. Starch solution was added to indicate end-point achievement until an abrupt color change. Room temperature oxygen nonstoichiometric (δ) was then calculated according to the consumption of thiosulfate solution. Thermo-gravimetric analysis (TGA) of STF powder was investigated in DTA/TA (NETZSH, STA 449 F3). The oxygen nonstoichiometric between 550 and 750oC was calculated with the following equation:

    δ=M×m0(M15.9994×δ0)×m15.9994×m0.

    where m0 represents starting weight, m represents actual weight, M represents STF molar mass of stoichiometric form, and δ0 represents room temperature oxygen nonstoichiometric, calculated from the iodometric titration.

    For conductivity testing, a bar-shaped STF was prepared by dry-pressing and then calcinated at 1250°C for 5 h. Four-probe DC configuration was applied for electrical conductivity (Keithley 2420 Source Meter). STF, SrFeO3–δ, SrNb0.1Fe0.9O3–δ, and SDC powders were pressed into pellets and pre-sintered. Netzsch dilatometer (DIL 402C/3/G) was exploited for TEC values of these samples from 200 to 1000°C.

    ASR values of STF/STF + SDC symmetrical cells were evaluated by Solartron electrochemical workstation (1287 + 1260A). The current–voltage–power (IVP) performances of SOFCs were measured via Keithley 2420 SourceMeter.

    Fig. 1(a) shows XRD patterns of STF and SrFeO3–δ oxide. At the Miller Index (1 1 0), the SrFeO3–δ sample exhibits a small peak beside the main peak, indicating that SrFeO3–δ perovskite consists of brownmillerite and cubic. It has been reported that the brownmillerite phase could converted to the cubic phase, when the temperature exceeds 900°C [21]. The corresponding Rietveld refinement result (Fig. 1(b)) reflects that the STF sample displayed cubic phase single perovskite (~100wt%), which has a space group of Pmˉ3m and lattice parameter a = b = c = 3.88732(4) Å. The fitting parameters of Rwp = 9.48%, Rp = 6.58%, and χ2 = 0.6576 (Rwp is weighted profile R-factor, Rp is expected R-factor, and χ is goodness of fitting) confirm the reliability of the Rietveld refinement, indicating a reliable fitting. EDS mapping images and HAADF-STEM demonstrate a homogeneous distribution of all the constituent elements of Sr, Ta, Fe, and O, as shown in Fig. 1(c)–(d). Above results revealed that 10mol% Ta5+ is successfully doped into the perovskite primitive cell. HR-TEM image and corresponding selected area electron diffraction patterns of STF (Fig. 1(e)–(f)) further confirm the cubic structure, and the lattice fringes of 0.287 nm along with the zone axes [1ˉ1ˉ1].

    Fig. 1.  (a) XRD patterns of synthesized SrFeO3–δ and STF powders, (b) Rietveld refinement profile of the STF powders at room temperature using XRD data, (c) EDS results of the STF powder, (d) HADDF-STEM image and corresponding EDS analysis of STF powder, (e–f) HR-TEM image and corresponding FFT of the STF sample along the [1¯1¯1] direction, respectively (inset of Fig. 1(e) was the magnified image).

    To further investigated the comprehensive characterization of the high-valent doped SF, Fig. 2 depicts X-ray photoelectron spectra (XPS) analysis results of SrFeO3–δ, SNF, and STF. Fig. 2(a) given the O 1s spectra of SrFeO3–δ, SNF, and STF. The four fitted peaks listed from high to low binding energy, attributed to adsorbed H2O, adsorbed oxygen (OH/O2), highly oxidative oxygen (O22/O), and the lattice oxygen (O2−), respectively. Table 1 shows peak deconvolution results of O 1s spectra, the percentages of lattice oxygen of SrFeO3–δ, SNF, and STF are 15.6%, 24.7%, and 18.6%, respectively. Fig. 2(b) and Table 2 display peak deconvolution results of Fe 2p3/2 spectra, the average valence state of Fe elements in SrFeO3–δ, SNF, and STF are 3.333, 3.394, and 3.280. The XPS results illustrated that there are two different charge compensation methods of the Nb5+ and Ta5+ doping in SrFeO3–δ. It was found that there was a dramatical increase of lattice oxygen ratio of SNF, while there was a decrease of the average valence state of Fe elements of STF. Considering the similar ionic radius, high-valent, and doping level, these two charge compensation methods could be ascribed to the different Pauling electronegativity of Ta5+ (1.80) and Nb5+ (1.87).

    Fig. 2.  XPS spectra of (a) O 1s and (b) Fe 2p of SrFeO3–δ, SNF, and STF.
    Table  1.  Peak deconvolution results of O 1s spectra for SrFeO3δ, SNF, and STF
    SampleO2–O22/OOH/O2H2O
    Binding energy / eVContent / %Binding energy / eVContent / %Binding energy / eVContent / %Binding energy / eVContent / %
    SrFeO3–δ528.115.6529.125.7531.351.4533.07.3
    SNF528.324.7529.324.4531.542.5533.18.4
    STF528.318.6529.424.9531.650.9533.15.6
     | Show Table
    DownLoad: CSV
    Table  2.  Peak deconvolution results of Fe 2p3/2 spectra for SrFeO3δ, SNF, and STF
    SampleFe4+Fe3+Fe2+
    Binding energy / eVContent / %Binding energy / eVContent / %Binding energy / eVContent / %
    SrFeO3–δ712.046.7710.539.9709.513.4
    SNF712.156.4710.626.6709.517.0
    STF712.346.9710.634.2709.418.9
     | Show Table
    DownLoad: CSV

    To study the thermal-induced characteristics of STF, electrical conductivity, O2-TPD, TGA, iodometric titration, and thermal expansion measurements were conducted. Fig. 3(a) shows the electrical conductivity of SrFeO3–δ, SNF, and STF. There is an inflection around 420°C, which is mainly associated with the thermal reduction of Fe4+, and is accompanied by the reduction of charge carrier at elevated temperatures. It can be seen that below 420°C, the electrical conductivity is increased with increasing temperature, whereas upon 420°C, the electrical conductivity is decreased with increasing temperature. At this temperature, the maximum conductivity achieves 21 S·cm–1. The electrical conductivity of STF during the working temperature range from 550 to 750°C is 10–17 S·cm–1. It is noted that the electrical conductivity of SNF and STF are smaller than SrFeO3–δ, illustrating the introduction of high valent Nb5+ and Ta5+ into the B-site of SrFeO3–δ restricted the electron transfer occurs through the B–O–B bond. Furthermore, the conductivity of STF is lower than SNF within the same temperature range, that is probably related to Nb5+ with higher electronegative will draw electron density from neighbouring Fe, and the average valence of Fe in SNF is higher than STF.

    Fig. 3.  (a) Electrical conductivity, (b) O2-TPD curve, (c) TGA curves, (d) oxygen nonstoichiometry (δ), (e) average valence state of iron ions (n), and (f) thermal expansion behavior of SrFeO3δ, SNF, STF, and SDC.

    The oxygen desorption onset temperature of STF occurred at 420°C (Fig. 3(b)), which is excellent in consistent with the inflection shown for STF electrical conductivity as a function of temperature. This result indicates that the reduction of transition metal (Fe4+ to Fe3+) in STF occurred above 420°C, which is higher than the onset temperature for SNF (~350°C) [16]. It is obvious that oxygen desorption peaked at 518°C, and the oxygen continued to be released even above the highest examination temperature of 930°C. When compared with Nb5+, the higher onset temperature results reveal that the doping of Ta5+ restricted the reduction of iron ions.

    Fig. 3(c) given TGA curves for SrFeO3–δ, SNF, and STF, and weight loss of these samples are 2.01wt%, 1.54wt%, and 0.86wt%, respectively. In the TGA curve for STF, the weight loss was initialized at about 420°C, which agrees with both the onset temperature of oxygen desorption and the electrical conductivity inflection. The release of oxygen from these samples along with the generation of extra oxygen vacancies. The iron ions undergo thermal reduction based on the local electrical neutrality principle. Assuming the interactions among the oxygen vacancies are negligible, the oxygen nonstoichiometry (δ) values can be taken as a measurement of the actual oxygen vacancy concentration. Partial replacement of iron ions in SrFeO3–δ with high valence ions of Ta5+ influences the actual Fe3+/Fe4+ ratio and oxygen vacancy concentration. The room temperature oxygen nonstoichiometric (δ) and iron average valence states (n) of STF, SNF, and SrFeO3–δ are listed in Table 3, as measured by iodometric titration. Compared to the δ and n of SNF, STF demonstrates higher oxygen vacancy concentration and lower average Fe valence. The room temperature δ values of SrFeO3–δ, STF, and SNF determined by iodometric titration were 0.268, 0.246, and 0.181, respectively. It shows that the Nb5+ doping of SrFeO3–δ could trigger a massive decrease in oxygen vacancies. Otherwise, the Ta5+ with the same high valence state exhibits completely different behaviour. The room temperature n values of SrFeO3–δ, STF, and SNF determined by iodometric titration were 3.464, 3.486, and 3.343, respectively, which demonstrates similar trends of the XPS fitting results. These results suggested that, compared with Nb5+ doping, the charge compensation of Ta5+ doping is mainly achieved by reducing Fe oxidation state rather than sacrificing the oxygen vacancy concentration. Temperature-dependent δ and average valence states of iron ions are shown in Fig. 3(d)–(e). At the SOFC working temperatures of 450–750°C, the δ value of STF increases from 0.262 to 0.331 and, accordingly, the state of iron ions decreases from 3.307 to 3.156, suggesting a dramatic increase in the content of oxygen vacancy concentration and a relatively moderate reduction in Fe4+ at higher temperatures. Besides, the behavior of TGA and iodometric titration are consistent with the conductivity trends, the lower electronic conductivity could be attributed to the restricted charge carriers.

    Table  3.  Average valence of Fe and oxygen nonstoichiometry for STF, SNF, and SrFeO3δ calculated from the iodometric titration
    Sample δ n Ref.
    STF 0.246 3.343 This work
    SNF 0.181 3.486 [16]
    SrFeO3–δ 0.268 3.464 This work
     | Show Table
    DownLoad: CSV

    The TEC of the cathode material should be matched to that of the electrolyte to avoid delamination during the thermal cycle. The thermal expansion behaviour of STF was evaluated, as shown in Fig. 3(f), the TEC values at the selected temperature ranges are presented in Table 4. Compared with SrFeO3–δ and SNF, the curve of STF shows an inflection at approximately 400°C, which is associated with the valence state change of iron ions as proved by O2-TPD results. The average TECs at the temperature range from 200 to 1000ºC are 34.1 × 10–6, 22.1 × 10–6, 14.6 × 10–6, and 13.3 × 10–6 K–1 for SrFeO3–δ, SNF, STF, and SDC, respectively. The TEC of the STF sample is significantly lower and quite nearly the SDC electrolyte [2122]. This suggests that the substitution of Ta5+ restricted the thermal reduction of iron in SrFeO3–δ, thus enhancing thermal compatibility with the electrolyte. In addition, the TECs of STF cathode show a good match with SDC electrolyte, which is optimized for the commercial utilization of solid oxide fuel cells.

    Table  4.  Average TEC values of SrFeO3δ, SNF, STF, and SDC at selected temperature ranges 10–6 K–1
    Temperature / °C SrFeO3–δ SNF STF SDC
    200–400 26.7 15.6 13.6 14.1
    400–1000 39.6 27.5 14.7 12.3
    200–1000 34.1 22.1 14.6 13.3
     | Show Table
    DownLoad: CSV

    The ORR activity of STF cathode was measured by electrochemical impedance spectra (EIS) examination. To investigate the relationship between electrode sintering temperature and electrochemical performance, the STF cathode was sintered at 800, 900, and 1000°C, respectively. ASR value was calculated from the intercept between high frequency and low frequency, which stands for ORR activity, as shown in Fig. 4(a). Fig. 4(b) displays the Nyquist plots of STF electrodes at 650°C with different cathode sintering temperatures, and fitting results are listed in Table 5. The fitting lines were well matched to the experimental data. It is revealed that higher sintering temperatures resulted in the enlargement of ASR values. The ASRs of the STF cathode sintered at 800°C are 0.015, 0.028, 0.067, 0.152, and 0.376 Ω·cm2 from 750 to 550°C at an interval of 50°C, which are only 28%–35% of the values obtained for the SrNb0.1Fe0.9O3–δ cathode. These values achieve the lowest ASRs among cobalt-free cathodes and exceed even those of some high-performance cobalt-containing cathodes. The activation energy (Ea) for STF is 113 kJ·mol–1, which is superior to the SrNb0.1Fe0.9O3–δ cathode (132 kJ·mol–1). This superior ORR activity of STF air electrode is mainly ascribed to the extremely high oxygen vacancy concentration.

    Fig. 4.  (a) Arrhenius plots of ASR for STF sintered at 800, 900, 1000, and STF + SDC (70:30, mass ratio) sintered at 800°C, (b) EIS for symmetrical cells with the configuration of STF |SDC| STF at 650°C with different cathode sintering temperatures, (c) XRD patterns of the STF with SDC electrolyte at different sintering temperatures, (d) ASR of SrFeO3δ, SNF, STF, and STF + SDC at 550, 600, 650, 700, and 750°C, respectively, (e) EIS for symmetrical cells SrFeO3δ, SNF, STF, and STF + SDC at 650°C, respectively, and (f) DRT plots of STF and STF + SDC (70:30, mass ratio) electrode at 650°C (γ(t) represents distribution of relaxation time).
    Table  5.  Fitting parameters for different cathodes at 650°C
    Sample R1 / (Ω·cm2) CPE1 / (F·cm2) R2 / (Ω·cm2) CPE2 / (F·cm2) Rtotal / (Ω·cm2)
    STF-800°C 0.023 0.95 0.043 0.87 0.065
    STF-900°C 0.026 0.93 0.094 0.92 0.120
    STF-1000°C 0.034 0.90 0.156 0.88 0.190
    STF + SDC-800°C 0.018 0.92 0.026 0.90 0.044
     | Show Table
    DownLoad: CSV

    The chemical stability between the STF and SDC electrolyte was demonstrated. The STF and SDC powders were mixed with a 1:1 mass ratio via high energy ball milling at 400 r/min for 1 h, then co-fired at 800, 900, and 1000°C for 2 h, respectively. XRD measurement was applied to detect the phase composition of STF, SDC, STF + SDC-800°C, STF + SDC-900°C, and STF + SDC-1000°C. XRD results shown in Fig. 4(c) manifested there is no any detectable new phase formation, suggesting the satisfactory chemical stability between STF and SDC electrolyte.

    By mixing primary material with the second phase to form composite electrode is usually used to improve the performance of air electrode. In order to further improve the ORR activity of the STF cathode, 30wt% SDC electrolyte powder was mixed with STF via high-energy ball milling. STF + SDC composite cathodes were painted onto two layers of SDC electrolyte, then calcinated at 800°C for 2 h.

    The ASRs of STF + SDC (70:30, mass ratio) composite cathode, between 550 and 750°C are shown in Fig. 4(d). It was found that the introduction of SDC electrolyte material improved the ORR activity, especially at low temperatures. The ASRs of the composite cathode are 0.012, 0.019, 0.044, 0.115, and 0.233 Ω·cm2 from 750 to 550°C at an interval of 50°C. The ASR of the composite cathode at 550°C is even slightly lower than for the benchmark cobalt cathode material (BSCF), which is 0.24 Ω·cm2 at the same temperature [20]. The Arrhenius plot in Fig. 4(a) demonstrates that the activation energy (Ea) for the STF + SDC (70:30, mass ratio) composite cathode is 108 kJ·mol–1, which is better than the pure STF cathode (113 kJ·mol–1) and SrNb0.1Fe0.9O3–δ cathode (132 kJ·mol–1). Fig. 4(e) displays the Nyquist plots of pure SF, SNF, STF electrode, and STF + SDC composite electrode at 650°C. The high ORR activity of the composite cathode may be due to the SDC electrolyte material extending the active oxygen reduction sites by increasing oxygen ionic conduction, therefore greatly reducing the cathode polarisation at lower working temperatures. The oxygen reduction reaction process of STF and STF + SDC electrodes was analysed by the distribution of relaxation time (DRT) technique. Based on the electrochemical impedance spectroscopy, DRT curves can be categorized as high-frequency (HF) portion (>104 Hz) which corresponds to charge transfer at triple phase boundaries (TPBs), the intermediate frequency (IF) portion (10–104 Hz) related to the surface exchange and ion diffusion in the air electrode, and low-frequency (LF) portion (10–2–10 Hz), associated with the gas diffusion [23]. Fig. 4(f) shows that HF and IF sections of resistances are significantly reduced through SDC introduction. The above result suggested that the addition of SDC electrolyte material efficiently improves surface exchange and bulk diffusion of oxygen ion, transformation in surface, and extension length of TPBs enhance the charge transfer between STF and SDC electrolyte.

    Fig. 5(a) shows I–V–P plots of the single fuel cell STF |SDC (~20 μm)| SDC–Ni (40:60, mass ratio), and pure hydrogen was taken as fuel at a flow rate of 60 mL·min–1. The peak power densities (PPDs) of fuel cells with STF cathode are 891, 749, 511, 285, and 136 mW·cm−2 from 650 to 450°C at an interval of 50°C. Open circuit voltages (OCVs) at recorded temperatures all exceed 0.8 V, represents that the tested SOFCs were intact and well-sealed. I–V–P properties of composite electrode were investigated under the cell structure STF + SDC (70:30, mass ratio) |SDC (~20 μm)| SDC–Ni (40:60, mass ratio). Fig. 5(b) shows the PPDs of SOFC with the STF + SDC composite cathode are 1117, 907, 635, 400, and 212 mW·cm−2 from 650 to 450°C at an interval of 50°C. The higher cell performance achieved by the 30wt% SDC composite cathode might be due to the favorable electrochemical characteristics, resulting from the optimized TPBs. Fig. 5(c)–(d) shows cross-section SEM images of the tested cell with pure STF and STF + SDC composite electrode, respectively. The SDC electrolytes have similar thickness, around 20 μm. The interfacial contacts between air electrode and electrolyte are satisfactory. It can be seen that STF electrode and STF + SDC composite electrode layers are tightly attached to SDC electrolyte, with no any detectable delimitations, verified the thermal compatibility.

    Fig. 5.  I–V–P plots and SEM images of SOFCs with configurations: (a, c) Ni + SDC |SDC| STF; (b, d) Ni + SDC |SDC| STF + SDC.

    In summary, STF cubic phase single perovskite oxide was synthesized via a solid-state method assisted with high-energy ball milling. The charge compensation of Ta5+ doping is mainly achieved by increasing the Fe3+/Fe4+ ratio rather than sacrificing the oxygen vacancy concentration. The average TEC of STF was 14.6 × 10–6 K–1 between 200 and 1000°C, which is quite near that of SDC electrolyte. STF cathode reached relatively low ASR of 0.015, 0.028, 0.067, 0.152, and 0.376 Ω·cm2 at 750, 700, 650, 600, and 550°C, respectively. With SDC as electrolyte and STF as cathode, the fuel cell provided PPDs of 891, 745, 510, 283, and 134 mW·cm–2 at 650, 600, 550, 500, and 450°C, respectively. To further improve the ORR activity of the STF cathode, 30wt% SDC powder was introduced, and the ASRs of STF + SDC composite cathode were 0.012, 0.019, 0.044, 0.115, and 0.233 Ω·cm2 at 750, 700, 650, 600, and 550°C, respectively. By introducing 30wt% SDC, the PPD of fuel cells with STF and STF + SDC composite cathode increased from 891 to 1117 mW·cm–2 at 650°C.

    This research was financially supported by the Natural Science Foundation of the Jiangsu Higher Education Institutions of China (No. 2018ND133J), the National Natural Science Foundation of China (Nos. 22309067 and 22101150), and the Natural Science Foundation of Jiangsu Province, China (No. BK20190965).

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

  • [1]
    Y. Zhang, B. Chen, D.Q. Guan, et al., Thermal-expansion offset for high-performance fuel cell cathodes, Nature, 591(2021), p. 246. DOI: 10.1038/s41586-021-03264-1
    [2]
    G.M. Yang, C. Su, H.G. Shi, et al., Toward reducing the operation temperature of solid oxide fuel cells: Our past 15 years of efforts in cathode development, Energy Fuels, 34(2020), No. 12, p. 15169. DOI: 10.1021/acs.energyfuels.0c01887
    [3]
    M. Wang, C. Su, Z.H. Zhu, H. Wang, and L. Ge, Composite cathodes for protonic ceramic fuel cells: Rationales and materials, Composites Part B, 238(2022), art. No. 109881. DOI: 10.1016/j.compositesb.2022.109881
    [4]
    J. Song, Y.Y. Birdja, D. Pant, Z.Y. Chen, and J. Vaes, Recent progress in the structure optimization and development of proton-conducting electrolyte materials for low-temperature solid oxide cells, Int. J. Miner. Metall. Mater., 29(2022), No. 4, p. 848. DOI: 10.1007/s12613-022-2447-y
    [5]
    W.N.A.W. Yusoff, N.A. Baharuddin, M.R. Somalu, A. Muchtar, N.P. Brandon, and H.Q. Fan, Recent advances and influencing parameters in developing electrode materials for symmetrical solid oxide fuel cells, Int. J. Miner. Metall. Mater., 30(2023), No. 10, p. 1933. DOI: 10.1007/s12613-023-2694-6
    [6]
    G.Y. Liu, F.G. Hou, S.L. Peng, X.D. Wang, and B.Z. Fang, Process and challenges of stainless steel based bipolar plates for proton exchange membrane fuel cells, Int. J. Miner. Metall. Mater., 29(2022), No. 5, p. 1099. DOI: 10.1007/s12613-022-2485-5
    [7]
    X. Yang, Z.H. Du, Q. Zhang, et al., Effects of operating conditions on the performance degradation and anode microstructure evolution of anode-supported solid oxide fuel cells, Int. J. Miner. Metall. Mater., 30(2023), No. 6, p. 1181. DOI: 10.1007/s12613-023-2616-7
    [8]
    P. Bhupaijit, C. Kaewsai, T. Suriwong, et al., Effect of Co2+ substitution in B-sites of the perovskite system on the phase formation, microstructure, electrical and magnetic properties of Bi0.5(Na0.68K0.22Li0.10)0.5TiO3 ceramics, Int. J. Miner. Metall. Mater., 29(2022), No. 9, p. 1798. DOI: 10.1007/s12613-021-2345-8
    [9]
    J. Yu, Z. Li, T. Liu, et al., Morphology control and electronic tailoring of Co xA y (A = P, S, Se) electrocatalysts for water splitting, Chem. Eng. J., 460(2023), art. No. 141674. DOI: 10.1016/j.cej.2023.141674
    [10]
    J.X. Pan, Y.J. Ye, M.Z. Zhou, et al., Improving the activity and stability of Ni-based electrodes for solid oxide cells through surface engineering: Recent progress and future perspectives, Mater. Rep. Energy, 1(2021), No. 2, art. No. 100025.
    [11]
    F.L. Liang, W. Zhou, and Z.H. Zhu, A highly stable and active hybrid cathode for low-temperature solid oxide fuel cells, ChemElectroChem, 1(2014), No. 10, p. 1627. DOI: 10.1002/celc.201402143
    [12]
    S.Y. Li, Z. Lü, B. Wei, et al., A study of (Ba0.5Sr0.5)1− x Sm xCo0.8Fe0.2O3− δ as a cathode material for IT-SOFCs, J. Alloys Compd., 426(2006), No. 1-2, p. 408. DOI: 10.1016/j.jallcom.2006.02.040
    [13]
    S. Park, S. Choi, J. Shin, and G. Kim, Tradeoff optimization of electrochemical performance and thermal expansion for Co-based cathode material for intermediate-temperature solid oxide fuel cells, Electrochim. Acta, 125(2014), p. 683. DOI: 10.1016/j.electacta.2014.01.112
    [14]
    Q. Huang, S.S. Jiang, Y.J. Wang, et al., Highly active and durable triple conducting composite air electrode for low-temperature protonic ceramic fuel cells, Nano Res., 16(2023), No. 7, p. 9280. DOI: 10.1007/s12274-023-5531-3
    [15]
    A. Wedig, R. Merkle, B. Stuhlhofer, H.U. Habermeier, J. Maier, and E. Heifets, Fast oxygen exchange kinetics of pore-free Bi1− xSr xFeO3− δ thin films, Phys. Chem. Chem. Phys., 13(2011), No. 37, p. 16530. DOI: 10.1039/c1cp21684h
    [16]
    S.S. Jiang, J. Sunarso, W. Zhou, J. Shen, R. Ran, and Z.P. Shao, Cobalt-free SrNb xFe1− xO3− δ (x = 0.05, 0.1 and 0.2) perovskite cathodes for intermediate temperature solid oxide fuel cells, J. Power Sources, 298(2015), p. 209. DOI: 10.1016/j.jpowsour.2015.08.063
    [17]
    Y.C. Zhou, X. Meng, X.J. Liu, et al., Novel architectured metal-supported solid oxide fuel cells with Mo-doped SrFeO3− δ electrocatalysts, J. Power Sources, 267(2014), p. 148. DOI: 10.1016/j.jpowsour.2014.04.157
    [18]
    J.H. Piao, K.N. Sun, N.Q. Zhang, X.B. Chen, S. Xu, and D.R. Zhou, Preparation and characterization of Pr1− xSr xFeO3 cathode material for intermediate temperature solid oxide fuel cells, J. Power Sources, 172(2007), No. 2, p. 633. DOI: 10.1016/j.jpowsour.2007.05.023
    [19]
    Y.F. Song, Y.B. Chen, M.G. Xu, et al., A cobalt-free multi-phase nanocomposite as near-ideal cathode of intermediate-temperature solid oxide fuel cells developed by smart self-assembly, Adv. Mater., 32(2020), No. 8, art. No. e1906979. DOI: 10.1002/adma.201906979
    [20]
    H.G. Shi, C. Su, X.M. Xu, et al., Building Ruddlesden-Popper and single perovskite nanocomposites: A new strategy to develop high-performance cathode for protonic ceramic fuel cells, Small, 17(2021), No. 35, art. No. e2101872. DOI: 10.1002/smll.202101872
    [21]
    F.L. Liang, Z.X. Wang, Z.R. Wang, J.K. Mao, and J. Sunarso, Electrochemical performance of cobalt-free Nb and Ta co-doped perovskite cathodes for intermediate-temperature solid oxide fuel cells, ChemElectroChem, 4(2017), No. 9, p. 2366. DOI: 10.1002/celc.201700236
    [22]
    D.M. Huan, L. Zhang, K. Zhu, et al., Oxygen vacancy-engineered cobalt-free Ruddlesden-Popper cathode with excellent CO2 tolerance for solid oxide fuel cells, J. Power Sources, 497(2021), art. No. 229872. DOI: 10.1016/j.jpowsour.2021.229872
    [23]
    T.H. Wan, M. Saccoccio, C. Chen, and F. Ciucci, Influence of the discretization methods on the distribution of relaxation times deconvolution: Implementing radial basis functions with DRT tools, Electrochim. Acta, 184(2015), p. 483. DOI: 10.1016/j.electacta.2015.09.097
  • Related Articles

    [1]Kevin Huang. A thermodynamic perspective on electrode poisoning in solid oxide fuel cells [J]. International Journal of Minerals, Metallurgy and Materials, 2024, 31(6): 1449-1455. DOI: 10.1007/s12613-023-2783-6
    [2]Shuo Han, Tao Wei, Sijia Wang, Yanlong Zhu, Xingtong Guo, Liang He, Xiongzhuang Li, Qing Huang, Daifen Chen. Recent progresses in the development of tubular segmented-in-series solid oxide fuel cells: Experimental and numerical study [J]. International Journal of Minerals, Metallurgy and Materials, 2024, 31(3): 427-442. DOI: 10.1007/s12613-023-2771-x
    [3]Shixue Liu, Zhijing Liu, Shuxing Zhang, Hao Wu. Lattice Boltzmann simulation study of anode degradation in solid oxide fuel cells during the initial aging process [J]. International Journal of Minerals, Metallurgy and Materials, 2024, 31(2): 405-411. DOI: 10.1007/s12613-023-2692-8
    [4]Wan Nor Anasuhah Wan Yusoff, Nurul Akidah Baharuddin, Mahendra Rao Somalu, Andanastuti Muchtar, Nigel P. Brandon, Huiqing Fan. Recent advances and influencing parameters in developing electrode materials for symmetrical solid oxide fuel cells [J]. International Journal of Minerals, Metallurgy and Materials, 2023, 30(10): 1933-1956. DOI: 10.1007/s12613-023-2694-6
    [5]Fuyuan Liang, Jiaran Yang, Haiqing Wang, Junwei Wu. Fabrication of Gd2O3-doped CeO2 thin films through DC reactive sputtering and their application in solid oxide fuel cells [J]. International Journal of Minerals, Metallurgy and Materials, 2023, 30(6): 1190-1197. DOI: 10.1007/s12613-023-2620-y
    [6]Xin Yang, Zhihong Du, Qian Zhang, Zewei Lyu, Shixue Liu, Zhijing Liu, Minfang Han, Hailei Zhao. Effects of operating conditions on the performance degradation and anode microstructure evolution of anode-supported solid oxide fuel cells [J]. International Journal of Minerals, Metallurgy and Materials, 2023, 30(6): 1181-1189. DOI: 10.1007/s12613-023-2616-7
    [7]Lifan Wang, Jingyue Wang, Leiying Wang, Mingjun Zhang, Rui Wang, Chun Zhan. A critical review on nickel-based cathodes in rechargeable batteries [J]. International Journal of Minerals, Metallurgy and Materials, 2022, 29(5): 925-941. DOI: 10.1007/s12613-022-2446-z
    [8]Jia Song, Yuvraj Y. Birdja, Deepak Pant, Zhiyuan Chen, Jan Vaes. Recent progress in the structure optimization and development of proton-conducting electrolyte materials for low-temperature solid oxide cells [J]. International Journal of Minerals, Metallurgy and Materials, 2022, 29(4): 848-869. DOI: 10.1007/s12613-022-2447-y
    [9]Zhi-yuan Chen, Liu-zhen Bian, Li-jun Wang, Zi-you Yu, Hai-lei Zhao, Fu-shen Li, Kuo-chih Chou. Topography, structure, and formation kinetic mechanism of carbon deposited onto nickel in the temperature range from 400 to 850℃ [J]. International Journal of Minerals, Metallurgy and Materials, 2017, 24(5): 574-583. DOI: 10.1007/s12613-017-1439-9
    [10]Bangwu Liu, Yue Zhang. Status and prospects of intermediate temperature solid oxide fuel cells [J]. International Journal of Minerals, Metallurgy and Materials, 2008, 15(1): 84-90. DOI: 10.1016/S1005-8850(08)60017-1
  • Cited by

    Periodical cited type(4)

    1. Hao Qiu, Jing Zhao, Zhixian Liang, et al. Chlorine turned the cobalt-free oxygen electrode of SrTa0.1Fe0.9O3–δ perovskite oxide towards protonic ceramic fuel cells. International Journal of Hydrogen Energy, 2025, 98: 353. DOI:10.1016/j.ijhydene.2024.12.090
    2. Hao Qiu, Mingzhuang Liang, Jing Zhao, et al. A-site cation deficient SrTa0.1Fe0.9O3-δ as a bi-functional cathode for both oxygen ion- and proton-conducting solid oxide fuel cells. Ceramics International, 2024, 50(20): 40500. DOI:10.1016/j.ceramint.2024.06.160
    3. Yuxuan Li, Yang Li, Shanshan Jiang, et al. La-doped Sr4Fe4Co2O13-δ as a promising in-situ self-assembled composite cathode for protonic ceramic fuel cells. Composites Part B: Engineering, 2024, 280: 111517. DOI:10.1016/j.compositesb.2024.111517
    4. Jie Yu, Guangming Yang, Zheng Li, et al. Boosting the lattice oxygen reactivity of perovskite electrocatalyst via less Ru substitution. International Journal of Hydrogen Energy, 2024, 84: 650. DOI:10.1016/j.ijhydene.2024.08.195

    Other cited types(0)

Catalog

    Figures(5)  /  Tables(5)

    Share Article

    Article Metrics

    Article views (650) PDF downloads (58) Cited by(4)

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return