Zhihong
Du
a,
Hailei
Zhao
*ac,
Yongna
Shen
a,
Lu
Wang
a,
Mengya
Fang
a,
Konrad
Świerczek
b and
Kun
Zheng
b
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
bAGH University of Science and Technology, Faculty of Energy and Fuels, Department of Hydrogen Energy, al. A. Mickiewicza 30, 30-059 Krakow, Poland
cBeijing Key Lab of New Energy Material and Technology, Beijing 100083, China
First published on 16th April 2014
Perovskites La0.3Sr0.7Ti1−xCoxO3 (LSTCs, x = 0.3–0.6) are systematically evaluated as potential cathode materials for solid oxide fuel cells. The effects of Co substitution for Ti on structural characteristics, thermal expansion coefficients (TECs), electrical conductivity, and electrochemical performance are investigated. All of the synthesized LSTCs exhibit a cubic structure. With Rietveld refinement on the high-temperature X-ray diffraction data, the TECs of LSTCs are calculated to be 20–26 × 10−6 K−1. LSTC shows good thermal cycling stability and is chemically compatible with the LSGM electrolyte below 1250 °C. The substitution of Co for Ti increases significantly the electrical conductivity of LSTC. The role of doping on the conduction behavior is discussed based on defect chemistry theory and first principles calculation. The electrochemical performances of LSTC are remarkably improved with Co substitution. The area specific resistance of sample La0.3Sr0.7Ti0.4Co0.6O3 on the La0.8Sr0.2Ga0.8Mg0.2O3−δ (LSGM) electrolyte in symmetrical cells is 0.0145, 0.0233, 0.0409, 0.0930 Ω cm2 at 850, 800, 750 and 700 °C, respectively, and the maximum power density of the LSGM electrolyte (400 μm)-supported single cell with the Ni–GDC anode, LDC buffer layer and LSTC cathode reaches 464.5, 648, and 775 mW cm−2 at 850 °C for x = 0.3, 0.45, and 0.6, respectively. All these results suggest that LSTC are promising candidate cathode materials for SOFCs.
The cobalt-based perovskites have attracted increasing attention as alternative cathode materials for IT-SOFCs due to their mixed-conducting characteristics and high electrocatalytic activity towards oxygen reduction.15–17 However, their practical application in IT-SOFCs is severely limited because of their large thermal expansion coefficient (TEC). The previous study reveals that introducing a stable Ti4+ ion to replace part of Co ions of Ba0.6Sr0.4CoO3−δ can successfully decrease the TEC.18 Recently, La0.5Sr0.5Co0.5Ti0.5O3 is reported to be a good candidate as a symmetrical electrode in IT-SOFCs. It shows good catalytic activity towards oxygen reduction in the cathode side and hydrogen oxidation in the anode side, and keeps good structural stability in both oxidizing and reducing atmospheres.19
La-doped SrTiO3 has been reported as a potential anode material for SOFCs due to its high electronic conductivity in reducing atmospheres and good chemical and structural stability upon redox cycling.20–23 Substituting Ti with Co for La0.3Sr0.7TiO3 (LST) could improve the oxygen ionic conductivity.24 Taking into account that Co ions usually possess high redox ability, more Co substitution for Ti has the potential to enhance the catalytic activity of LST towards oxygen reduction, and thus can turn LST into a cathode material. The stable oxidation state of Ti ions in air ensures good structural stability and a lower TEC value for LST. As symmetrical electrode materials, La0.5Sr0.5Co0.5Ti0.5O3,19 La0.4Sr0.6Ti1−yCoyO3−δ (ref. 25) and La2−xSrxCoTiO6,26 have been investigated recently. These studies focus mainly on the lattice structure evolution of (LaSr)(TiCo)O3 with chemical composition and environmental atmosphere, but less on the electrochemical performance, except for the work on La0.5Sr0.5Co0.5Ti0.5O3.
As B-site elements, the different electronic structures of Ti and Co ions endow perovskite oxides with much different properties, in terms of structural stability, lattice defect, electronic conductivity, and catalytic activity. The Co/Ti ratio will have a strong impact on the electrode performance of (LaSr)(TiCo)O3 materials. In this work, La0.3Sr0.7Ti1−xCoxO3 (LSTC, x = 0.3–0.6) materials are prepared and characterized as cathode materials, with the aim to get a deep understanding of the effect of Co content on the crystal structure, electrical conductivity and electrochemical properties. First principles calculation is employed to elucidate the electronic conduction behavior of LSTC.
The obtained final solution was heated in a water bath at 80 °C until a gel was formed. The gel was transferred into an oven and heated at 250 °C to get a fluffy precursor, which was ground and subsequently calcined at 800 °C for 6 h with an interval at 400 °C for 2 h to obtain LSTC powders. The prepared LSTC powders were uniaxially pressed into pellets (19 mm in diameter) and rectangular bars (40 mm × 7 mm × 3 mm) with an appropriate amount of polyvinylalcohol (PVA, 1 wt%) as a binder, followed by sintering at 1200 °C (x = 0.3, 0.45) and 1150 °C (x = 0.6) for 10 h in air to get dense samples for electrical conductivity measurement.
For symmetrical cells of LSTC|LSGM|LSTC, LSTC slurries were screen-printed on both sides of the LSGM electrolyte symmetrically (active area 0.5 cm2), followed by calcining at 1200 °C for 2 h. For a single cell with the configuration of Ni–GDC|LDC|LSGM|LSTC, the LDC slurry was deposited on one side of the LSGM electrolyte (active area 0.78 cm2) and fired at 1400 °C for 2 h. Subsequently, the Ni–GDC anode slurry was screen-printed on the LDC layer and fired at 1300 °C for 2 h. Finally LSTC (x = 0.3–0.6) was applied on the other side of the LSGM electrolyte and calcined at 1200 °C for 2 h. The active area of both anode and cathode was 0.5 cm2. Ag paste was used as the current collector, which was painted in a grid structure on both sides of the cells and fired at 650 °C for 0.5 h. The final cells were sealed on an alumina tube with a ceramic-based sealant (Cerama-bond 552-VFG, Aremco). Humidified H2 (∼3% H2O) was fed as fuel to the anode with a flow rate of 40 ml min−1, and pure O2 or air (100 ml min−1) as an oxidant to the cathode.
![]() | ||
Fig. 1 XRD patterns of La0.3Sr0.7Ti1−xCoxO3 samples sintered for 10 h at 1200 °C for x = 0.3 and 0.45 and at 1150 °C for x = 0.6. |
In order to characterize the structure stability of LSTC during the heating process, the samples are subjected to HT-XRD examination. The typical HT-XRD results for x = 0.6 are shown in Fig. 2. The results reveal that the synthesized LSTC (x = 0.3–0.6) keep their cubic structure over the temperature range of RT–900 °C, and no structure change or phase segregation is detected, indicating the good structural stability of the synthesized LSTC samples. To obtain detailed information about the crystal structure of the synthesized LSTC at different temperatures, Rietveld refinement is performed on the HT-XRD data using the GSAS/EXPGUI program. For all the LSTC samples, the Pmm cubic perovskite structure is employed as an initial model, where La/Sr is located at 1a (0, 0, 0) site, Ti/Co at 1b (1/2, 1/2, 1/2) site, and O at 3c (0, 1/2, 1/2) site. The typical refinement results for samples with x = 0.45 at 25 and 900 °C are illustrated in Fig. 3. The refinement shows good agreement between the calculated and observed profiles. The refined structure parameters for samples at 25 and 900 °C are summarized in Table 1. The lattice parameter of samples decreases with increasing content of Co for both at 25 and 900 °C, which is consistent with the variation of the diffraction peaks in Fig. 1. This implies that the content of Co4+ ions increases in LSTC, taking into account the ionic radius of Co4+ (0.67 Å), Co3+ (LS: 0.685 Å, HS: 0.75 Å) and Ti4+ (0.745 Å),30 and the fact that the Co3+ ion tends to take a higher spin state at high temperatures.31–33
![]() | ||
Fig. 3 XRD patterns and Rietveld refinement results of the sample with x = 0.45 at 25 °C (a) and 900 °C (b): observed (cross symbols) and calculated (continuous line). |
x = 0.3 | x = 0.45 | x = 0.60 | ||||
---|---|---|---|---|---|---|
25 °C | 900 °C | 25 °C | 900 °C | 25 °C | 900 °C | |
a = b = c (Å) | 3.8783(22) | 3.9451(6) | 3.8667(18) | 3.9391(19) | 3.8591(9) | 3.9366(9) |
La/Sr Uiso | 0.0150(2) | 0.0356(2) | 0.0150(2) | 0.0348(3) | 0.0176(1) | 0.0361(2) |
Ti/Co Uiso | 0.0077(3) | 0.0206(3) | 0.0074(3) | 0.0182(4) | 0.0105(3) | 0.0281(4) |
O Uiso | 0.0257(6) | 0.0458(7) | 0.0217(6) | 0.0511(17) | 0.0291(7) | 0.0677(9) |
χ 2 | 2.408 | 3.251 | 2.479 | 2.919 | 4.010 | 4.814 |
R p (%) | 1.75 | 1.83 | 1.62 | 1.56 | 1.49 | 1.39 |
R wp (%) | 2.46 | 2.83 | 2.35 | 2.50 | 2.42 | 2.56 |
With HT-XRD data, the lattice parameter of samples at different temperatures can be calculated by Rietveld refinement. The obtained lattice parameter variations of the investigated samples LSTC at different temperatures are depicted in Fig. 4. With the fitted linear slope, the thermal expansion coefficients of samples are derived, which are shown in the inset of Fig. 4. Two slopes with a bending at 300 °C can be observed for all three samples. The increased slope in the high temperature range, corresponding to a larger TEC value, is associated with the so-called chemical expansion, which is caused by the loss of lattice oxygen, reduction of B-site ions and/or transition from low-spin to high-spin state of partial Co ions.31,34,35 In a low temperature range (RT–300 °C), all samples show almost the same slope. However, in a high temperature range, the slope increases significantly with the Co content, corresponding to the increased TEC. The TEC is 20.7(7), 23.6(3) and 26.3(2) × 10−6 K−1 for the LSTC samples with x = 0.3, 0.45 and 0.6, respectively. The increased TEC with increasing Co content is related to the weak Co–O bond compared to the Ti–O bond. The high content of Co in the sample will induce more lattice oxygen loss and thus more oxygen vacancy generation and more Co ion reduction, which thus result in large lattice expansion.
![]() | ||
Fig. 4 Temperature dependence of the refined lattice parameter variations of different samples. The inset is the calculated thermal expansion coefficients. |
This assumption is supported by TG results. As shown in Fig. 5, there is a slow weight loss in the range RT–300 °C, corresponding to a mass change of ca. 0.09%, which probably originated from the evaporation of absorbed water. Above 300 °C, a significant weight loss is observed for the sample with x = 0.6, while only a slight weight loss is detected for the sample with x = 0.3, demonstrating the large lattice oxygen loss of the sample with a high Co content. This is consistent with the TEC results. Although a high Co content leads to a high TEC value, it can also give rise to more oxygen vacancies, which are beneficial to the electrode reaction process. It is worth noting that the TEC derived from the lattice parameter variation upon temperature is different from that obtained by directly measuring the dense sample. Considering that the dense sample usually contains more or less pores, the calculated TEC from lattice parameter change is higher than the practically measured value.17 Nevertheless, the calculated TEC of LSTC (x = 0.3–0.6) prepared in this work, 20–26 × 10−6 K−1, is lower than the tested TEC of La0.3Sr0.7CoO3−δ (28.8 × 10−6 K−1), SrCo0.8Ti0.2O3−δ (28.3 × 10−6 K−1)36 and BSCF (24.9–27.3 × 10−6 K−1).35,37
![]() | (1) |
e′ + h˙ → nil | (2) |
At 800 °C, the conductivity of LSTC reaches 24 and 110 S cm−1 for x = 0.45 and 0.6, respectively, which are comparable to or even higher than that of BSCF.35
For charge compensation of La0.3Sr0.7TiO3 without Co-doping, it is believed that the excessive positive charge produced by La-doping at Sr-site is balanced by the formation of A-site vacancies.25,38 With respect to the system of La0.3Sr0.7Ti1−xCoxO3, when x = 0.3, the excessive positive charge of La can be balanced by the substitution of Co for Ti, as expressed in eqn (3), provided that the oxygen vacancy concentration is limited in air. Because both Co3+ and Ti4+ ions are in the stable oxidation state and no electron defect is generated, the sample La0.3Sr0.7Ti0.7Co0.3O3 shows a much lower conductivity.
![]() | (3) |
For further substitution of Co for Ti (x = 0.45 and 0.6), the extra Co ions will take the oxidation state of Co4+ in order to keep the neutrality of the material. Considering that considerable oxygen vacancies may produce in Co-rich samples, which will cause the generation of free electrons as shown in eqn (1), the solid solution formula of La0.3Sr0.7Ti1−xCoxO3−δ can be written as:
La0.33+Sr0.72+Ti1−x4+Co0.3+2δ3+Cox−0.3−2δ4+O3−δ |
Accordingly, the concentration [Co4+], corresponding to the concentration of electron–holes [h˙], is proportional to the Co content x. As a result, the electronic conductivity increases with the Co content in LSTC materials, which is consistent with the results shown in Fig. 7(b).
To confirm the assumption discussed above, XPS examination is performed to identify the oxidation state of Co ions in LSTC. The evolution of the Co 2p photoelectron spectra as a function of doping amount is shown in Fig. 8. Two broad peaks, belonging to Co 2p3/2 and Co 2p1/2 electrons, were observed, both of which can be deconvoluted into two peaks, assignable to Co3+ and Co4+, respectively.39–41 With the integrated area of the peaks, the contents of each oxidation state of Co ions are calculated, which are listed in Fig. 8. Besides Co3+, a considerable amount of Co4+ is detected in both samples and the content of Co4+ increases with increasing Co doping level, from 24% (x = 0.45) to 48% (x = 0.6). This result is consistent with the assumption of charge compensation. A satellite peak at around 786 eV was evident, suggesting a small amount of Co2+ coexisting on the surface of LSTC particles.40,42
In order to evaluate the performance stability of LSTC during the thermal cycling process, the conductivity of the sample with high Co content (x = 0.6) is tested under different cycles. The result is shown in Fig. 7(d). The sample shows similar conductivity during the second and third cycles but much lower than that in the first cycle. The conductivity of mixed conductors depends strongly on the thermal history they experienced. As mentioned above, some lattice oxygen may lose at high temperature in the sintering process, which can lead to the decrease in conductivity. During the cooling process, these oxygen vacancies can be incorporated into the bulk of the sample again in order to maintain the thermodynamic equilibrium. In the case of fast cooling, however, this process cannot be thoroughly completed due to kinetic reasons, and more oxygen vacancies will remain, leading to reduced electrical conductivity. Because the thermal insulation system of the furnace for sample preparation in our lab is much better than that for electrical conductivity measurement, the sample, which has ever experienced the conductivity test and thus been subjected to the fast cooling process, usually presents a much lower conductivity value than the initial one. Fortunately, the sample displays similar electrical conductivity values in the subsequent cycles, demonstrating the highly reversibility of the structure and the good thermal cycling performance of LSTC as the cathode for SOFCs.
To get insight into the effect of Co-doping on the electronic conductivity of the LSTC, first-principles calculations are performed to get the density of states (DOS) based on density functional theory using Materials Studio software. For simplicity, La2Sr6Ti8O24 and La2Sr6Ti4Co4O24 are approximately used as the calculation models for La0.3Sr0.7TiO3 with and without Co-doping (Fig. 9). The calculated results are shown in Fig. 10. Significantly, Co-doping changes the density of states of electrons in LSTC. LSTC has transformed from n-type conductor (LST) to p-type conductor, since the Fermi level shifts from the conduction band (Fig. 10(a)) to the valence band (Fig. 10(b)). This finding is in good agreement with the experimental results (Fig. 7) and the previous reports that LST exhibits n-type conduction behavior under reducing atmosphere while Sr-doped LaCoO3 shows p-type conduction characteristics.43–45 The remarkably high conductivity values of LSTC (x = 0.45 and 0.6) are mainly attributed to the narrow bandgap and the observed strong hybridization of the Co and O states at the valence band edge and conduction band (Fig. 10(b) and (c)). Both of them lead to low activation energy for electron–holes jumping along the Co–O–Co bonding network.
![]() | ||
Fig. 10 Density of states of (a) La2Sr6Ti8O24 and (b) La2Sr6Ti4Co4O24; (c) detailed view of the states of O p, Co d and Ti d. Dashed lines represent the Fermi energy. |
T/°C | 700 | 750 | 800 | 850 | |
---|---|---|---|---|---|
0.3 | R/Ω cm2 | 0.498 | 0.268 | 0.162 | 0.104 |
C/F cm−2 | 0.073 | 0.061 | 0.055 | 0.052 | |
F/Hz | 4.38 | 9.76 | 18.01 | 29.36 | |
ASR/Ω cm2 | 0.498 | 0.268 | 0.162 | 0.104 | |
0.45 | R H/Ω cm2 | 0.0195 | 0.0090 | 0.0045 | 0.0023 |
C H/F cm−2 | 0.167 | 0.160 | 0.111 | 0.116 | |
F H/Hz | 48.91 | 109.97 | 317.34 | 612.19 | |
R L/Ω cm2 | 0.220 | 0.091 | 0.053 | 0.035 | |
C L/F cm−2 | 0.202 | 0.211 | 0.191 | 0.168 | |
F L/Hz | 3.57 | 8.27 | 15.73 | 26.84 | |
ASR/Ω cm2 | 0.2395 | 0.1000 | 0.0575 | 0.0373 | |
0.6 | R H/Ω cm2 | 0.0150 | 0.0079 | 0.0043 | 0.0025 |
C H/F cm−2 | 0.102 | 0.086 | 0.085 | 0.101 | |
F H/Hz | 102.44 | 236.01 | 430.75 | 637.36 | |
R L/Ω cm2 | 0.078 | 0.033 | 0.019 | 0.012 | |
C L/F cm−2 | 0.575 | 0.475 | 0.393 | 0.339 | |
F L/Hz | 3.56 | 10.08 | 21.30 | 37.66 | |
ASR/Ω cm2 | 0.0930 | 0.0409 | 0.0233 | 0.0145 |
It is suggested that the high frequency arc is mainly related to the charge transfer process, while the low frequency arc is associated with the molecular oxygen dissociation process,46–49 as expressed in eqn (4) and (5). Since the low frequency arc is much larger than the high frequency arc, the rate-limiting step of the electrode reaction should be the molecular oxygen dissociation processes.
Oad + e′ ↔ O′ad | (4) |
O2,ad ↔ 2Oad | (5) |
The various resistances versus reciprocal temperature, accompanying with activation energy, are shown in Fig. 12. All the resistances (RH and RL) decrease noticeably with increasing temperature, indicative of thermal activation behavior of the electrode reaction process. With increasing Co content, all the RH and RL decrease significantly, demonstrating that Co-doping improves the molecular oxygen dissociation and charge transfer processes of the electrode. The sample with x = 0.6 yields the lowest area-specific resistance of 0.0145 Ω cm2 at 850 °C, followed by x = 0.45 and 0.3 with ASR of 0.0373 and 0.104 Ω cm2, respectively, indicating that the electrode performance is significantly enhanced by Co substitution. The ASR of the LSTC electrode with x = 0.6 is superior to and comparable with the reported typical cathode materials based on the LSGM electrolyte, as summarized in Table 3.
![]() | ||
Fig. 12 Polarization resistances versus reciprocal temperature for the La0.3Sr0.7Ti1−xCoxO3−δ cathode obtained in air. |
Composition | Temperature/°C | ASR/Ω cm2 | Reference |
---|---|---|---|
SmBaCo2O5+x | 800 | 0.031 | 50 |
PrBa0.5Sr0.5Co2O5+x | 800 | 0.027 | 51 |
GdBaCo2O5+δ | 800 | ∼0.138 | 52 |
Ba0.5Sr0.5Co0.8Fe0.2O3−δ | 800 | ∼0.075 | 52 |
Sr0.7Y0.3CoO2.65−δ | 800 | 0.11 | 53 |
Sm0.5Sr0.5CoO3−δ | 800 | 1.34 | 54 |
La0.6Sr0.4Fe0.8Co0.2O3−δ | 700 | 0.1 | 55 |
Pr2NiO4+δ | 700 | 0.23 | 55 |
La1.7Ca0.3Ni0.7Cu0.3O4+δ | 800 | 0.099 | 56 |
SrCo0.9Nb0.1O3−δ | 800 | 0.029 | 57 |
BaCo0.7Fe0.2Nb0.1O3−δ | 750 | 0.06 | 58 |
Ba0.9Co0.7Fe0.2Nb0.1O3−δ | 800 | 0.02 | 59 |
700 | 0.0930 | This work | |
La0.3Sr0.7Ti0.4Co0.6O3−δ | 750 | 0.0409 | This work |
800 | 0.0233 | This work |
In order to evaluate the cathodic properties of LSTC, the LSGM electrolyte-supported single cell with the configuration of LSTC/LSGM/LDC/(Ni–GDC) is constructed and the performance is examined (Fig. 13). The open circuit voltage (OCV) of the cells is about 1.10 V at 850 °C, which is very close to the theoretical value (∼1.13 V), indicating that the gas leakage is very small. The cell performance of LSTC is remarkably enhanced by Co substitution. At 850 °C, the maximum power density with O2 as an oxidant is 464.5, 648 and 775 mW cm−2 for x = 0.3, 0.45 and 0.6, respectively, revealing the LSTC as a quite potential cathode material. With air as an oxidant, the cell with LSTC (x = 0.45) delivers the maximum power density of 597 mW cm−2 at 850 °C (Fig. 14), just slightly lower than that in pure O2. Considering that the thickness of the LSGM electrolyte used in the single-cell is about 400 μm, the cell performance is acceptable, which can be further enhanced by using a thinner LSGM electrolyte and optimizing the electrode structure. The cell microstructure after the test is provided in Fig. 15. The electrodes still maintain the porous structure and good connection with the LSGM electrolyte. The excellent performance of electrolyte-supported single cells with the LSTC cathode is a strong indication of its potential as a cathode material for SOFCs.
![]() | ||
Fig. 14 Voltage and power density versus current density plots for single-cell LSTC (x = 0.45)/LSGM/LDC/(Ni–GDC) with humidified pure hydrogen as fuel and air as an oxidant. |
![]() | ||
Fig. 15 SEM micrographs of the cross-section of the tested cells: LSTC cathode on LSGM electrolyte with (a) x = 0.3, (b) x = 0.45, (c) x = 0.6; (d) Ni–GDC anode with an LDC buffer layer. |
This journal is © The Royal Society of Chemistry 2014 |