Weicheng
Feng‡
ab,
Geng
Zou‡
ab,
Tianfu
Liu‡
ab,
Rongtan
Li‡
ab,
Jingcheng
Yu
ab,
Yige
Guo
ab,
Qingxue
Liu
ab,
Xiaomin
Zhang
a,
Junhu
Wang
ac,
Na
Ta
a,
Mingrun
Li
a,
Peng
Zhang
d,
Xingzhong
Cao
d,
Runsheng
Yu
d,
Yuefeng
Song
*ae,
Meilin
Liu
*e,
Guoxiong
Wang
*af and
Xinhe
Bao
a
aState Key Laboratory of Catalysis, Dalian National Laboratory for Clean Energy, iChEM (Collaborative Innovation Center of Chemistry for Energy Materials), Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian, 116023, China. E-mail: wanggx@dicp.ac.cn
bCollege of Energy, University of Chinese Academy of Sciences, Beijing, 100049, China
cCenter for Advanced Mössbauer Spectroscopy, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian, 116023, Liaoning, China
dMulti-disciplinary Research Division, Institute of High Energy Physics, Chinese Academy of Sciences, Beijing, 100049, P. R. China
eSchool of Materials Science and Engineering, Georgia Institute of Technology, Atlanta, 30332, USA
fDepartment of Chemistry, Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, iChEM (Collaborative Innovation Center of Chemistry for Energy Materials), Fudan University, Shanghai, 200438, China
First published on 22nd January 2025
While the effects of Sr segregation on the performance and stability of perovskite electrodes in solid oxide electrolysis cells (SOECs) have been widely studied, most attention has been focused on surface Sr segregates, with the impact of the resulting Sr deficiencies within the bulk phase of the electrodes largely ignored. Here, we report our findings from an investigation into the impact of Sr deficiencies in the SrCo0.7Fe0.3O3−δ (SCF) lattice and surface Sr segregates on the electrochemical behavior of well-controlled anode materials. Results demonstrate that Sr deficiencies in the perovskite lattice significantly enhance bulk oxygen ion transport, while surface Sr segregates suppress oxygen vacancy formation at interfaces, resulting in a reduced rate of oxygen exchange and lower surface electrical conductivity. Our study provides critical insights into the roles of bulk Sr deficiencies and surface Sr segregates, particularly their effects on oxygen vacancy formation, electrical conductivity, oxygen ion transport, and the overall rate of a high-temperature oxygen evolution reaction.
Broader contextSolid oxide electrolysis cells (SOECs) represent a promising technology to efficiently convert renewable energy into stable chemical energy. However, Sr segregation onto the anode surface is generally considered as a common stability issue for a high-temperature oxygen evolution reaction (OER). Most studies have primarily focused on the influence of surface Sr segregates, while the effect of bulk Sr deficiencies on the elementary processes of electrode reactions has been rarely investigated. Herein, density functional theory calculations combined with various physicochemical characterization studies and electrochemical impedance spectra confirm that the generated bulk Sr deficiencies in the SrCo0.7Fe0.3O3−δ anode during Sr segregation can lift the O 2p band center toward the Fermi level and promote the formation of oxygen vacancies, thus accelerating the bulk oxygen ion transportation and the electron transfer processes in the bulk, while the surface Sr segregates decelerate the oxygen exchange and electron transfer processes. This work provides a comprehensive understanding of the Sr segregation process during a high-temperature OER and is instructive for the knowledge-based design of advanced anode materials with enhanced performance and stability. |
Perovskite oxides with mixed electron–ion conductivity are commonly used as anode materials due to their high conductivities and oxygen vacancy concentrations, which are essential for high-temperature electrocatalytic reactions.8–11 Numerous studies have confirmed that replacing A-site cations (of an oxidation state +3 or higher) with Sr2+ increases the concentration of oxygen vacancies, thus improving the OER activity.12–16 However, Sr segregation onto the electrode surface is commonly observed under typical SOEC operating conditions (at high temperatures and high oxygen partial pressures) due to cation size mismatch between doped Sr2+ and the host cation along with the electrostatic interactions between the dopant and the surrounding lattice ions.17–20 Recently, Sr segregation and its effect on the performance degradation of perovskite electrodes have been extensively studied.6 Yildiz et al. revealed that the enlarged band gap with Sr segregation on the SrTi1−xFexO3 surface inhibited the electron transfer process from the bulk to surface adsorbed oxygen species and decreased the active oxygen vacancy concentration.18 Rupp et al. observed the spontaneous formation of Sr-rich composite oxides on a LSC thin film at high temperatures and confirmed their adverse effects on oxygen surface exchange kinetics.21
However, contradictory effects of Sr segregation on the electrochemical processes have also been reported. Li et al. discovered that the presence of SrO on the La0.6Sr0.4CoO3−δ surface could prevent intrinsic Sr segregation and improve the electrochemical performance efficiently at 600 °C.22 Additionally, many Sr-containing perovskite anodes also demonstrated good stability during long-term operation.23–26 For example, Barnett et al. achieved a stability of more than 1000 h for the SrTi0.3Fe0.7O3−δ anode at high current density, with only a slight increase in resistance.27 Our previous work demonstrated satisfactory stability of the SrCo0.7Fe0.3O3−δ anode for 487 h,16 with a slight decrease in the SOEC performance accompanied by a significant amount of Sr segregates on the anode surface. In fact, Sr deficiencies are inevitably generated in the bulk lattice of the electrode as Sr segregates to the surface. However, most studies have primarily focused on the influence of surface Sr segregates, while the effect of bulk Sr deficiencies on the elementary processes of electrode reactions has been rarely investigated. As the anodic OER involves not only the surface electron transfer and oxygen ion exchange but also the transport of oxygen ions and electrons through the bulk phase of the electrode,4,28,29 taking into consideration the effects of both surface Sr segregates and bulk Sr deficiencies is of great significance to assess the real effects of Sr segregation on the high-temperature OER performance and stability.
In this study, we carried out various ex situ and in situ physicochemical characterization studies, density functional theory (DFT) calculations, and electrochemical impedance spectroscopy (EIS) analysis to explore the influence of Sr segregation on the OER performance and stability of the SrCo0.7Fe0.3O3−δ (SCF) anode from two aspects: surface Sr segregates and bulk Sr deficiencies. DFT calculation results show that the bulk Sr deficiencies lift the O 2p band center and promote oxygen vacancy formation, which is inhibited by surface Sr segregates. Physicochemical measurements confirm the promoted bulk oxygen ion mobility and electron transfer by the bulk Sr deficiencies and the decelerated surface oxygen exchange and electron transfer processes by surface Sr-enriched oxides. These results reveal the dual-impact mechanism of the Sr segregation process and explain the absence of significant performance degradation during long-term operation in actual electrode systems, which is instructive for the design of stable and efficient anode materials for SOECs.
![]() | ||
Fig. 2 XPS spectra of (a) Fe 2p and (b) Co 2p of SCF and SCF-d. 57Fe Mossbaüer spectra of (c) SCF and (d) SCF-d samples. |
Moreover, 57Fe Mössbauer spectroscopy was used to characterize the changes in valence states and coordination environments of Fe species after Sr segregation. The results are shown in Fig. 2c and d, and Table S7 (ESI†). According to the variations in isomer shift (IS), magnetic field strength (MF), and quadrupole splitting (QS),38 the spectra of both SCF and SCF-d can be fitted in three doublets (labelled as D1, D2, and D3) and one singlet (labelled as S1) with their areas representing the proportion of different Fe species. The ISs of the D1 subspectra of SCF and SCF-d are 0.23(1) mm s−1 and 0.27(1) mm s−1, respectively, which are attributed to Fe3+ species with pyramidal coordination,39 and the D3 of SCF and SCF-d are 0.29(3) and 0.33(3) mm s−1, respectively, which are attributed to Fe3+ species with octahedral coordination.40 The D2 and S1 of both samples with ISs less than 0 are both attributed to Fe4+ species with octahedral coordination. The distinctions in peak shapes, IS and QS of D2 and S1 are mainly caused by differences in adjacent metal elements, as well as the symmetry and distortion of the coordination polyhedra.41–43 With the formation of Sr deficiencies, the proportion of Fe4+ species in SCF-d increases from 43.72% to 51.76%, and the average valence of Fe increases from +3.43 to +3.52.
The increase in the valence states of the B-site transition metal cations due to Sr deficiencies in SCF-d creates more Fe3+–Fe4+ and Co3+–Co4+ ion pairs, thereby expanding the electronic conduction pathway.44 This is reflected in the increased electrical conductivity of SCF-d (155.76 S cm−1) compared to that of SCF (113.55 S cm−1) at 800 °C (Fig. S4, ESI†). In contrast, the electrical conductivity of SCF-O2 decreases significantly to 50.47 S cm−1, primarily due to surface Sr segregates. Generally, the electronic conductivity in bulk materials is orders of magnitude higher than the oxygen ion conductivity.45–50 As the high-temperature OER involves both electronic conduction and oxygen ion conduction processes, oxygen ion transport at the interface plays a decisive role in determining the resistance and the final electrochemical performance.
For SCF and SCF-d, three types of oxygen atoms bonded with Co/O (S1), Co/Sr (S2) and Fe/Sr (S3) atoms on the surfaces are removed, while for SCF-O2, three types of oxygen atoms bonded with Sr/Fe (S1), Sr/Sr (S2) and Sr/Co (S3) atoms at the SrO/SCF-d interface are removed (Fig. S5, ESI†), and the required energy is defined as the surface oxygen vacancy formation energy (Evfs). The calculation results, presented in Fig. 3e and Table S8 (ESI†), show that, compared with SCF, Evfs at each position of SCF-d is reduced, while Evfs at the SrO/SCF-d interface significantly increases. This indicates that Sr deficiencies promote the formation of oxygen vacancies, whereas surface Sr segregates suppress the generation of interface oxygen vacancies.
Similarly, oxygen atoms at four different sites (B1–B4) are removed from the bulk models of SCF and SCF-d, and the required energy is defined as the bulk oxygen vacancy formation energy (Evfb). As shown in Fig. 3f and Fig. S6 (ESI†), and listed in Table S9 (ESI†), all Evfb values at four positions significantly decrease from 1.10, 1.00, 0.78, and 0.73 eV in SCF to −0.12, −0.12, −0.66, and −0.33 eV in SCF-d, indicating that Sr deficiencies in SCF-d promote the formation of oxygen vacancies in the bulk phase.
The oxygen surface exchange process is examined directly by 18O2 isotope exchange experiment, as shown in Fig. 4b and c and Fig. S7 (ESI†). It is widely acknowledged that the exchange between gaseous 18O2 molecules and lattice 16O ions in a perovskite oxide can be expressed as follows:57
18O2(g) + 16O(s) ↔ 18O(s) + 18O16O(g) | (1) |
18O2(g) + 16O–16O(s) ↔ 18O–18O(s) + 16O2(g) | (2) |
The exchange between one lattice 16O ion and an 18O2 molecule is reflected by the signal with m/z of 34, and the exchange of two lattice 16O ions with an 18O2 molecule is reflected by the signal with m/z of 32. The signal with m/z of 36 represents the consumption of 18O2 with increasing temperature. As shown in Fig. 4a–c, the desorption temperatures of the signals with m/z of 32 and 34 decrease from 327 °C and 340 °C for SCF to 271 °C and 286 °C for SCF-d with Sr deficiencies, but increase back to 304 °C and 320 °C for SCF-O2, while the total amount of 18O2 is simultaneously reduced (Fig. S6, ESI†). Therefore, it is reasonable to conclude that the Sr deficiencies induced in the bulk can enhance the transport and surface exchange of oxygen species. However, the enhancement is hindered by the accumulation of Sr-enriched phase on the surface. Moreover, electrical conductivity relaxation (ECR) profiles in Fig. S8 (ESI†) show that SCF-d reaches the equilibrium state faster than SCF at 800 °C as oxygen partial pressure is reduced from 0.21 to 0.05 atm.58–60 According to the fitting results of the ECR profiles, the oxygen chemical diffusion coefficient (Dchem) of SCF-d is 7.05 × 10−5 cm2 s−1, which is much higher than that of SCF (3.20 × 10−5 cm2 s−1), further confirming enhanced bulk oxygen ion mobility due to Sr deficiencies under typical SOEC operating conditions.
Furthermore, quasi in situ time-of-flight secondary ion mass spectrometry (TOF-SIMS) measurements are performed to investigate the effect of Sr deficiencies on the bulk oxygen ion mobility under the working conditions (Fig. S9, ESI†). The two-dimensional spectra of the 16O− and 18O− signals shown in Fig. 4d and e and Fig. S10 (ESI†) are obtained by bombarding the cross-section of the pretreated SOEC (see the Experimental section for details) with an ion beam and collecting the generated secondary ions. The 18O− signal intensity of the SCF anode is significantly higher than that of SCF-d. Since both cells are pretreated under the same conditions, equal amounts of 18O2− are transferred from the cathode to anode; therefore, the weaker intensity of the 18O− signal in the cross-section of the SCF-d anode indicates that more 18O2− in the SOEC with the SCF-d anode is transferred to the anode surface and evolves into a gas. Additionally, the two-dimensional spectra are integrated and normalized along the y-axis in Fig. 4f, and the 18O− signal at the cross-section of the SCF-d anode is significantly decreased compared with that of SCF, indicating that the bulk oxygen ion transport process in SCF-d is accelerated. Therefore, physicochemical measurements combined with DFT calculations above show that Sr deficiencies promote the formation of oxygen vacancies, resulting in enhanced oxygen ion mobility, while the surface Sr segregates do the opposite.
As the OER kinetics is closely related to the oxygen vacancy concentration and the oxygen ion mobility of the anode, high-temperature in situ electrochemical XPS measurements are performed to explore the impact of Sr deficiencies on the OER process. As shown in Fig. 4g–i, O 1s XPS spectra can be fitted into four peaks, corresponding to four different surface oxygen species.51,61–63 The peak with the lowest binding energy at approximately 528.40 eV belongs to Olat., and the other three oxygen species at 529.60, 531.00 and 532.50 eV represent highly active O− or O22−, adsorbed OH−, and adsorbed H2O, respectively; these three oxygen species are classified as Oads. species. With the temperature increasing to 1000 K, most Oads. species desorb from the anode surface, and the proportion of Oads. species is reduced by 60.17% in SCF-d–SDC and 57.83% in SCF-SDC, which indicates a higher oxygen vacancy concentration in the SCF-d–SDC anode and is consistent with the results of DFT calculations and O2-TPD. After applying a constant current of 1 mA to SOECs (Fig. 4i and Table S10, ESI†), the content of Oads increases by 23.26% in SCF-SDC and 40.05% in SCF-d–SDC because of the spillover processes of oxygen ions from the triple phase boundaries onto the anode surface under anodic polarization. The higher content of surface spillover oxygen species on SCF-d indicates that Sr deficiencies could accelerate the elementary process of the OER due to the enhanced oxygen ion mobility and surface diffusion capacity.
EIS spectra of the single cells with a LSCF-SDC cathode and different anodes are shown in Fig. 5c and Fig. S13 (ESI†). It is evident that the presence of Sr deficiencies could efficiently decrease RΩ from 0.32 Ω cm2 for the SCF-SDC anode to 0.28 Ω cm2 for the SCF-d–SDC anode with the concurrent decrease of Rp from 0.16 to 0.09 Ω cm2. However, after the formation of Sr segregates on the SCF-O2-SDC anode surface, Rp remains almost the same as that of the SCF-d–SDC anode, but RΩ increases from 0.28 to 0.33 Ω cm2, which are consistent with the EIS results of a symmetrical cell as shown in Fig. 5a and b.
DRT analysis of the impedance data is presented in Fig. 5d, providing more detailed information on the frequency dispersion of electrode processes.64 Three distinct peaks are identified in the DRT plots, corresponding to three electrode processes with different relaxation times. Typically, the process taking place at a high frequency range (P1) is attributed to the transportation of oxygen ions through the electrode/electrolyte, the process at an intermediate frequency range (P2) may be associated with electron transfer and surface reactions at the anode interface, and the process at a low frequency range (P3) may be related to the cathodic CO2 reduction processes, including CO2 adsorption, dissociation of carbonate species, and CO desorption.65–68 Moreover, the complex nonlinear least squares method (CNLS) fittings of EIS curves are performed based on the DRT results to further quantitatively analyze each elementary reaction process (Fig. 5a and Table S11, ESI†), and the equivalent circuit model used in the curve fitting is shown in Fig. S14 (ESI†). One resistor (RΩ) and three R–C parallel components (R1–C1 to R3–C3) are connected in series to simulate RΩ and the three electrode reaction processes corresponding to P1, P2, and P3. Among these three processes, P1 changes sensitively with anode materials. For example, the R1 value decreases by 44.33% as the anode is changed from SCF-SDC to SCF-d–SDC, due to the increased oxygen ion mobility and the accelerated rate of electron transport within the Sr-deficient SCF-d bulk phase. In contrast, the R1 value increases by 59.49% after the formation of surface Sr segregates, resulting from the blocking effect of the surface Sr-enriched oxide. The bulk electronic conductivity primarily determines the electron transport process within the material, corresponding to RΩ observed in the EIS. In contrast, the exchange and transport of oxygen ions at the interface significantly influence the anodic reaction processes, corresponding to the R1 process in the EIS and DRT. The electrochemical analysis results indicate a reduction in R1 for SCF-d, suggesting that Sr deficiencies enhance the performance of electrochemical reactions and promote overall kinetics by improving ion mobility and the interfacial charge transfer rate. Thus, the performance of the SOEC for CO2 electrolysis is enhanced (Fig. 5e) by the Sr deficiencies in the bulk phase but hindered by the Sr segregates on the surface (Fig. 5f).
SOECs were fabricated by the screen-printing method. First, anode slurries were printed onto one side of the LSGM pellet and then sintered in a muffle furnace at 1200 °C in air for 2 h. Afterward, the cathode slurry was printed onto the other side of the LSGM pellet and then sintered in a muffle furnace at 1100 °C in air for 2 h. The anode slurry was printed onto both sides of the LSGM pellets and then sintered in air at 1200 °C for 2 h in a muffle furnace to obtain ready-to-use symmetrical cells. The areas of the anode and cathode used for electrochemical measurements and quasi in situ time of flight secondary ion mass spectrometry (TOF-SIMS) measurements were 0.50 and 0.78 cm2, respectively, which were 0.28 and 0.50 cm2 for those used for in situ X-ray photoelectron spectroscopy (XPS) measurements. The thickness of the anode is about 40 μm (Fig. S15, ESI†)
Powder X-ray diffraction (XRD) experiments were carried out at different temperatures from RT to 800 °C in air using a PANalytical X’pert PPR diffractometer equipped with a Cu Kα radiation source (λ = 1.5418 Å) operating at 40 kV and 40 mA. The Rietveld refinement results were obtained using Highscore plus software.
X-ray photoelectron spectroscopy (XPS) measurements were conducted using an ESCALAB 250Xi spectrometer. Al Kα was used as the X-ray excitation source, and the vacuum in the analysis chamber was maintained at 7 × 10−5 Pa. The sample data were calibrated using the C 1s peak at a binding energy of 284.6 eV, and the data were analyzed and fitted using the XPS Peak software.
Positron annihilation tests were carried out at the Institute of High Energy Physics of the Chinese Academy of Sciences. The 22Na radioactive source was used as the positron source with a source intensity of approximately 13 μCi. Two identical samples with a thickness of 1 mm and a diameter of 13 mm were tightly sandwiched between the two sides of the 22Na source. A pair of BaF2 scintillator detectors were used to detect the gamma rays released after the positron annihilation, and the positron annihilation lifetime spectrum was measured by a fast-slow coincidence measurement technique with a total lifetime spectrum count of 2 million to ensure statistical accuracy. The time resolution of the positron annihilation lifetime spectrometer was approximately 210 ps, and the electronics of the measurement system were standard NIM modules from EG&G, USA.
Ex situ X-ray photoelectron spectroscopy (XPS) tests were performed using an ESCALAB 250Xi spectrometer. Al Kα was used as the X-ray excitation source, and the vacuum of the analytical chamber was 7 × 10−5 Pa. The C 1s peak with a binding energy of 284.6 eV was used to calibrate the energy, and the results were analyzed and fitted using XPSPeak software. In situ XPS spectra of all the samples were collected on an EnviroESCA-Near Ambient Pressure XPS. The O 1s signals were collected at RT and 800 °C, and then a constant current was applied to the SOECs, and the corresponding O 1s signal was collected.
Ar ion etching treatment was performed using a laboratory-built XPS equipped with a SPECS XR-50 X-ray gun and a SPECS PHOIBOS-100 energy analyzer. The vacuum of the analytical chamber during etching was 5 × 10−6 mbar, the etching voltage was 2.5 kV, and the etching depth was about 20–40 nm. XPS spectra of Sr 3d were collected after etching to observe the content of Sr species on the surface.
Mössbauer spectroscopy measurements were performed on a Topologic Systems MFD-500AV-02 Mössbauer spectrometer, and the results were analyzed and fitted by Mosswinn software.
O2-temperature programmed desorption (O2-TPD) experiments were performed using an Auto Chem II 2920 chemisorption instrument (Micromeritics Chemisorption Analyzer, USA). The samples were first pretreated at 800 °C in 5% O2–95% He flow for 1 h, and then the samples were cooled to room temperature and held for 30 min under He flow. After, the O2 desorption process was performed at a heating rate of 10 °C min−1 under He with a flow rate of 50 mL min−1 from 50 °C to 750 °C, the desorption products were detected using an OmniStar online quadrupole mass spectrometer, and the signal at m/z of 32 was detected for analysis. For the 18O2 isotope exchange tests, the samples were preheated at 800 °C in 20% O2–80% He for 1 h. Then, the sample was placed under He flow for 30 min. After that, the sample was cooled to room temperature in He flow. Then, 5% 18O2–95% He was introduced into the sample at a flow rate of 10 mL min−1 along with a heating rate of 10 °C min−1 from 50 to 750 °C. The signals at m/z of 32, 34 and 36 were detected for analysis.
TOF-SIMS measurements were carried out in TOF mode using SIMS 5-100 (IONTOF, Germany) at Ningbo Institute of Materials Technology & Engineering, Chinese Academy of Sciences. Typically, the cathode was fed with 5% 18O2–95% He, while a constant current of 200 mA was applied to the SOEC for 45 min. The cross-section of the treated SOEC was tested using a 30 kV Bi ion source to observe the imaging signals of 18O− and 16O− at the interface of the electrolyte and the anode.
57Fe Mössbauer spectrometry was carried out at 300 K for 24 hours. Data analysis and fitting were entrusted to the Mössbauer Spectroscopy Testing Center within the institute, using the MossWinn software.
Electrical conductivity relaxation (ECR) test was performed at temperatures of 600, 700, and 800 °C. The testing method involved introducing air into the test tube at a flow rate of 100 mL min−1 at constant temperature, while measuring the I–t curve at 1 mV. Once the curve stabilized, the atmosphere was rapidly switched to 5% O2–95% Ar at the same flow rate of 100 mL min−1. The test was concluded once the curve stabilized again.
The electrical conductivity test was conducted using a two-electrode four-terminal method. A constant voltage was applied using a PGSTAT302N Autolab electrochemical workstation, and the conductivity was calculated from the chronoamperometric curve. The test was performed in an air atmosphere at temperatures ranging from 300 to 800 °C.
The formation energy of oxygen vacancies was investigated by removing one of the four kinds of oxygen sites to simulate the oxygen vacancies, which are the two top sites and two of the four in-plane sites. For the calculations of oxygen vacancy formation energy (Evf), Evf was defined as follows:
Evf = (Evac + 1/2EO2) − Eclean |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ee05056h |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |