Next Article in Journal
A Sustainable Strategy for Co-Melting of Electroplating Sludge and Coal Gasification Slag: Metals Recovery and Vitrified Product Valorization
Previous Article in Journal
Design and Development of a Neural Network-Based End-Effector for Disease Detection in Plants with 7-DOF Robot Integration
Previous Article in Special Issue
Multi-Physics Coupling Simulation of H2O–CO2 Co-Electrolysis Using Flat Tubular Solid Oxide Electrolysis Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

CO2 Poisoning of Solid Oxide Fuel Cell Cathodes: Mechanisms, Solutions, and Progress

by
Fang Liu
1,
Quan Luo
2,
Meishen Sun
2,
Zhaoqi Song
3,*,
Junbiao Li
3,*,
Bin Chen
3,* and
Yuan Zhang
2,3,*
1
School of Dentistry, Shenzhen University Medical School, Shenzhen 518060, China
2
College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China
3
Institute of Deep Earth Sciences & Green Energy, Shenzhen University, Shenzhen 518060, China
*
Authors to whom correspondence should be addressed.
Processes 2025, 13(12), 3931; https://doi.org/10.3390/pr13123931 (registering DOI)
Submission received: 12 October 2025 / Revised: 2 November 2025 / Accepted: 22 November 2025 / Published: 5 December 2025
(This article belongs to the Special Issue Recent Advances in Fuel Cell Technology and Its Application Process)

Abstract

Conventional energy resources have been constrained by their inefficient utilization and present a severe impact on the human living environment, and there is an urgent need to develop energy technologies with high efficiency, low carbon emissions, and environmental cleanliness. Solid oxide fuel cells (SOFCs) have been recognized as a highly efficient and clean energy conversion device that directly converts chemical energy in fuels into electricity, holding promising prospects for addressing the issues of low efficiency and environmental concerns associated with conventional energy resources. However, under practical operation conditions, the cathodes of SOFCs are often exposed to various contaminations including working environment-induced degradation, cathode poisoning, and corrosion. This review summarizes the severe performance degradation of SOFC cathodes caused by CO2 poisoning, analyzes recent research findings on cathode durability under CO2-containing atmospheres, and provides an overview of the reported strategies for enhancing CO2 tolerance.

Graphical Abstract

1. Introduction

In recent years, growing concerns over environmental pollution and energy crisis issues have intensified the demands for sustainable solutions, as illustrated in Figure 1a [1,2]. Solid oxide fuel cells (SOFCs) have emerged as a promising power generation system owing to their ability for direct, efficient, and low-pollution energy conversion [3]. SOFCs find potential applications in diverse areas such as distributed generation, transportation, combined heat and power (CHP), and trigeneration systems (Figure 1b) [4]. This broad applicability comes from the fact that SOFCs are all-solid-state electrochemical energy conversion devices capable of directly and efficiently converting the chemical energy from various hydrocarbon fuels, including hydrogen, methane, and ethanol, to electricity [5]. However, under real-life operating conditions, the high temperatures for SOFC operation pose challenges, (Figure 1c) [6], such as compatibility issues among cell components (Figure 1d) [7], shortened operational lifespan (Figure 1e) [8], mismatched thermal expansion coefficients (Figure 1f) [9], cathode delamination, and degradation induced by the surrounding environment [10,11,12]. Consequently, exploring the fundamental mechanisms underlying the degradation of intermediate-temperature SOFCs (IT-SOFCs) is of particular importance.
SOFCs consist of anodes, electrolyte, and cathodes, and among them, the cathode plays a particularly crucial role that significantly determines the overall high-performance output of the cell. The performance is intrinsically dependent on the kinetics of oxygen reduction and transport occurring at its surface [13]. Additionally, the stability of the cathode material constitutes a core factor governing the long-term operational performance, service span, and commercialization of SOFCs. As a key cell component, the cathode operates under harsh high-temperature conditions (700–900 °C), catalyzing the oxygen reduction reaction (ORR) [14,15], with its stability directly impacting the energy conversion efficiency and system reliability. During practical operation, cathodes must continuously endure high-temperature oxidizing atmospheres, fluctuations in gas composition, and thermal cycling stresses induced by frequent start–stop operations. Insufficient stability leads to significant performance degradation or even failure, with the primary failure mechanisms categorized as chemical degradation and mechanical failure. Chemical degradation manifests as the segregation of active elements forming insulating phases. For instance, the intermediate-temperature cathode La0.6Sr0.4Co0.2Fe0.8O3−δ (LSCF), known for its excellent mixed conductivity, experiences performance decay during prolonged operation due to the formation of segregating phases like SrCO3 or SrCrO4. Similarly, cathodes with perovskite-related structures, such as (Ba,Sr)(Co,Fe)O3−δ (BSCF), (La,Sr)CoO3−δ (LSC), and the layered perovskite PrBa0.5Sr0.5Co0.5Fe0.5O5+δ (PBSCF) [16,17,18,19], display sharp breakdown due to similar segregation phenomena. Mechanical failures, on the other hand, originate from structural instability under thermal cycling stresses. Mismatched thermal expansion coefficients (TECs) readily induce interfacial delamination and crack formation. Zhang et al. developed the Srx(Yy(Nb0.1Co0.9)1−y)O3−δ (SYNC) cathode by combining a negative thermal expansion component, achieving synergistic enhancement in thermomechanical compatibility and high catalytic activity [20,21]. Additionally, a chemically induced hydration strategy was introduced to reinforce the mechanical strength of protonic ceramic fuel cell cathodes. An as-prepared BaCo0.7Ce0.15Y0.15O3−δ (s-BCC-Y) cathode exhibited strengthened grain intergranular bonding and an 86% increase in fracture strength after hydration treatment, maintaining structural integrity even after 35 thermal cycles between 300 and 600 °C [22]. Similarly, a Nb-doped BaCe0.1Fe0.8Nb0.1O3−δ (BCFNb) cathode demonstrates excellent mechanical stiffness along with robust ORR/oxygen evolution reaction (OER) activity [23].
It is noteworthy that during prolonged SOFC operation, the cathode is inevitably and continuously exposed to CO2 in ambient air. Such exposure leads to cathode material poisoning, triggering significant chemical degradation and mechanical failure, which severely impair the cell’s activity and durability [24,25]. This degradation mechanism is particularly pronounced in ABO3-type perovskite cathode materials, especially in those containing alkaline earth metal elements, which are structurally susceptible to CO2 poisoning [26]. For instance, the performance degradation of cathode materials such as Ba0.5Sr0.5Co0.8Fe0.2O3−δ (BSCF) and Pr0.6Sr0.4Co0.5Fe0.5O3−δ (PSCF) has been demonstrated to be closely associated with the presence of carbon dioxide [27,28]. Considering that practical SOFC systems demand stable operation for thousands of hours or even longer, it is essential to achieve the long-term stability of cathodes under the combined effects of thermal fields, chemical environments (particularly CO2), and mechanical stresses in this field. To address this challenge, researchers are consistently driven to develop various cathode material modification strategies, including techniques such as multi-step solution infiltration, forming composite cathodes with materials like gadolinium-doped ceria (GDC), and introducing isovalent metal ions for doping. These strategies aim to significantly enhance the cathode’s tolerance and long-term stability under harsh operating conditions. Therefore, assessing the poisoning behavior of a cathode and its underlying mechanisms is crucial for developing commercially viable SOFCs. This paper systematically evaluates the poisoning impact of CO2 on SOFC cathode performance.
Figure 1. (a). Schematic of an SOFC operating with hydrogen as the fuel. Reproduced with permission from Ref. [2]. (b). SOFC-absorption cooling building cogeneration system configuration. Reproduced with permission from Ref. [29]. (c). The possible poisoning and corrosions of cathode materials. (d) Delamination of SOFC cell components due to high operation temperature. Reproduced with permission from Refs. [7,9]. (e) Cracking behavior in SOFC electrolyte due to heating–cooling cycling, subfigure (a) displays an initial anode and its substrate as its co-fired state, subfigure (b) displays a reduced state, and (c) displays a re-oxidized state. Reproduced with permission from Refs. [8,9]. (f). Possible mechanism of cell cracking and delamination during cell operation at high temperature. Reproduced with permission from Ref. [9].
Figure 1. (a). Schematic of an SOFC operating with hydrogen as the fuel. Reproduced with permission from Ref. [2]. (b). SOFC-absorption cooling building cogeneration system configuration. Reproduced with permission from Ref. [29]. (c). The possible poisoning and corrosions of cathode materials. (d) Delamination of SOFC cell components due to high operation temperature. Reproduced with permission from Refs. [7,9]. (e) Cracking behavior in SOFC electrolyte due to heating–cooling cycling, subfigure (a) displays an initial anode and its substrate as its co-fired state, subfigure (b) displays a reduced state, and (c) displays a re-oxidized state. Reproduced with permission from Refs. [8,9]. (f). Possible mechanism of cell cracking and delamination during cell operation at high temperature. Reproduced with permission from Ref. [9].
Processes 13 03931 g001

2. CO2 Poisoning Mechanism

2.1. Description of the Basic CO2 Poisoning Process

The mechanism of CO2 poisoning in SOFC cathode materials primarily involves both physicochemical adsorption and structural transformations. Among these, the affinity of alkaline earth metal oxide (AEMO) surfaces for CO2 follows the trend: BaO > SrO > CaO > MgO [30]. Driven by physical adsorption, CO2 initially captured on the AEMO surface forms monolayer or multilayer coverage. This not only impedes oxygen adsorption and diffusion but also obstructs the cathode’s porous structure and covers active sites, ultimately inhibiting oxygen transport and reaction kinetics.
During chemical adsorption, CO2 readily reacts with alkaline elements (e.g., Ba2+ and Sr2+) present on the cathode surface to form stable carbonates (e.g., SrCO3 and BaCO3). The stability of these carbonates increases with the ionic radius of the alkaline earth metal (Figure 2a) [31,32]. Taking Ba0.5Sr0.5Co0.8Fe0.2O3−δ (BSCF), a classic high-performance mixed-conductor cathode material, as an example [33], it exhibits higher ionic conductivity and oxygen permeability compared to lanthanum-based perovskite materials due to the larger ionic radius of barium [34]. BSCF demonstrates exceptional oxygen reduction electrocatalytic activity, high electrochemical performance, superior oxygen permeation flux, and good phase stability at 600 °C. These properties enhance both the bulk oxygen diffusion rate and the surface exchange kinetics [35,36].
However, as illustrated in Figure 2b, research by Yan et al. [37] revealed that the BSCF cathode is extremely sensitive to CO2 within a temperature range of 450–750 °C, particularly under low-temperature operating conditions and even at very low CO2 concentrations (<1%). For instance, in an air atmosphere with a low content of 0.28% CO2, the open-circuit voltage and peak power density of BSCF decrease dramatically. When the CO2 concentration reaches 0.85%, the peak power density at 600 °C is nearly halved. Moreover, the performance degradation is even more severe and largely irreversible at 500 °C and 450 °C [38]. This clearly demonstrates that even relatively small amounts of CO2 in air cause significant and irreversible poisoning effects on BSCF. Yan et al. [39] further reported that after exposure to air at 600 °C for 7 days, the degradation rate of oxygen exchange kinetics in BSCF approached approximately 80%.
Furthermore, CO2 adsorption may induce structural transformations in cathode materials. For instance, Zhang et al.’s investigation of the BaFeO3−δ (BF) cathode revealed that CO2 exposure triggers in situ formation of barium carbonate (BaCO3) [40]. This subsequently drives a comprehensive structural reconstruction of the cathode, resulting in a core-shell architecture comprising a BaCO3 outer shell and a dual-phase core (BaFe2O4 and BaFeO3 phases). Similarly, Tan et al. reported that the exposure to CO2 induces surface strontium segregation and corrosion in La0.6Sr0.4Co0.2Fe0.8O3−δ (LSCF) cathodes [41]. Other prominent cathode materials, including (Ba,Sr)(Co,Fe)O3−δ (BSCF), (La,Sr)CoO3−δ (LSC), and PrBa0.5Sr0.5Co1.5Fe0.5O5+δ (PBSCF), also undergo severe degradation through elemental segregation mechanisms, leading to detrimental alterations in material structure [18].
Figure 2. (a). Ellingham diagram for the decomposition of carbonates under different CO2 partial pressures. Reproduced with permission from Ref. [42]. (b). Electrochemical impedance spectra measured at various times after CO2 was supplied at 450 °C. Redrawn from Ref. [37].
Figure 2. (a). Ellingham diagram for the decomposition of carbonates under different CO2 partial pressures. Reproduced with permission from Ref. [42]. (b). Electrochemical impedance spectra measured at various times after CO2 was supplied at 450 °C. Redrawn from Ref. [37].
Processes 13 03931 g002

2.2. General Impact Patterns of A-Site Elements on Poisoning

The crystal structure is crucial for the performance of SOFC cathodes, as it directly determines their electrocatalytic activity and long-term stability. Various oxides with distinct crystal structures, such as perovskite-type and double perovskite-type oxides, are widely demonstrated to act as superior SOFC cathode materials. Among these, perovskite-structured oxides are considered the most promising cathode materials due to their high catalytic activity toward the ORR. For these materials, the physicochemical properties (e.g., electrical conductivity) can be tuned over a wide range by substituting the A-site ions (typically La, Sr, Ba, etc.), achieving conductivities from nearly pure electronic to almost pure ionic. Representative high-performance perovskite cathodes include La1−xSrxCo1−yFeyO3−δ (LSCF), Ba1−xSrxCo1−yFeyO3−δ (BSCF), Sm0.5Sr0.5CoO3−δ (SSC), and PrBaCo2O6−δ (PBC) [43,44,45,46]. These materials typically contain abundant alkaline earth metal cations (e.g., Sr2+, Ba2+) occupying the A-site, which induce the formation of substantial oxygen vacancies, thereby significantly enhancing oxygen transport properties. Consequently, they are widely employed for their excellent power output and stability. However, these very A-site alkaline earth elements make the cathodes susceptible to significant performance degradation or even material failure when exposed to CO2, severely compromising cell durability [47,48].
As reported by Shao et al. [35], the barium-based material BSCF undergoes surface segregation under ambient air due to CO2 exposure, forming carbonate layers and secondary oxide phases that impair surface oxygen exchange kinetics through microstructural and conductivity changes [49,50]. Strontium-based cathode materials typically exhibit poor CO2 tolerance, since CO2 readily reacts with Sr to form SrCO3 [51,52], while CO2 absorption concurrently blocks active sites on the cathode surface [53]. Experimental and simulation studies by Khromushin et al. revealed significant interactions between CO2 and strontium-doped La1−xSrxMnO3−δ (LSM), as shown in Figure 3a. Their CO2 temperature-programmed desorption (TPD) spectra demonstrated that the amount of CO2 desorbed increased markedly with higher strontium doping levels and directly proved that the higher Sr content in LSM enhances its CO2 adsorption capacity [54]. This poisoning phenomenon, initiated by high-basicity elements at the A-site, is prevalent across a wide range of cathode materials, such as BaCoxFeyZr2O3−δ (BCFZ), SrCo0.9Nb0.1O3−δ (SCN), SrCo0.9Sc0.1O3−δ (SCS), BaCe0.9Y0.1O3−δ, Sr(CoFe)O3−δ, and BaPr0.7Gd0.3O3−δ [49,55,56,57,58,59,60]. Zhang et al. evaluated the CO2 tolerance of the Sr-based cathode SrSc0.175Nb0.025Co0.8O3−δ (SSNC) using electrochemical impedance spectroscopy (EIS) [61]. After exposure to 10 vol% CO2 for 15 min at 500 and 600 °C, followed by its removal, SSNC and BSCF exhibited an increase in area-specific resistance (ASR) by a magnitude of 20–45 times, whereas at 700 °C the relative increase was only 4–9 times. SSNC demonstrated superior CO2 tolerance compared to BSCF, attributed to the higher basicity of BaO versus SrO, which facilitates more facile carbonate formation, and the surface faced a sharp change, as illustrated in Figure 3b. As shown in Figure 3c, Hu et al. documented that SrCO3 formation on the surface of La0.6Sr0.4Co0.2Fe0.8O3−δ (LSCF) cathodes in 0–10% CO2-air mixtures causes initial performance degradation [62]. The fundamental reason lies in the high basicity of the alkaline earth metals, which significantly promotes the chemisorption of CO2; the resultant carbonate species or surface segregation layers cover ORR-active sites. Furthermore, the persistent accumulation of these carbonates can lead to irreversible structural damage.
Additionally, based on the Lewis acid-base theory, BaO exhibits the strongest basicity, surpassing that of SrO and La2O3. Correspondingly, oxides with stronger basicity react more readily with CO2. Furthermore, the Ellingham data presented in Figure 2a corroborate this trend, demonstrating that Ba2+ possesses a greater thermodynamic driving force for carbonate formation compared to Sr2+ and La3+ [63]. Therefore, collectively, the basicity of the A-site element is the key factor governing a cathode’s susceptibility to CO2 poisoning. The stronger the basicity, the higher the tendency for chemisorption reactions with CO2, leading to more severe poisoning effects such as carbonate formation or surface segregation.
To make the discussion quantitative, we consider the surface-carbonation equilibria of A-site oxides with CO2. The overall reaction can be described as follows (Equations (1) and (2)):
BaO (SrO, La2O3) + CO2(g) ⇌ BaCO3 (or SrCO3, La2O2CO3),
Δ G ° ( T ) + RT   ln   P C O 2   =   0 .
Here, R is the gas constant, and T is the temperature. The magnitude of ln P C O 2 reflects the extent of the carbonates forming reaction, and a value above 105 indicates complete reaction completion. Therefore, the CO2 resistance of different metal elements can be compared by evaluating their corresponding ln P C O 2 values.
At a given temperature, carbonate stability follows Equation (2). Carbonates are stable when the ambient P C O 2 lies above the boundary, and they decompose when it lies below. The stability boundaries follow the order BaCO3 < SrCO3 < La2O2CO3, indicating that Ba and Sr carbonates stabilize more readily than La oxy-carbonate. In LSC/LSCF, Sr surface segregation effectively lowers the threshold for SrCO3 formation, so even sub-percent CO2 at approximately 600–700 °C can sustain persistent surface coverage, whereas La-rich surfaces are less susceptible under the same conditions. Increasing the temperature and decreasing P C O 2 drive decarbonation, which helps maintain accessible oxygen-exchange sites and limits interfacial secondary-phase formation.

2.3. General Impact Patterns of B-Site Elements on Poisoning

The B-sites of mixed ionic–electronic conductor (MIEC) oxide cathodes are typically occupied by transition metal ions. Based on the dominant cation type, they are classified as Mn-based cathodes (e.g., La1−xSrxMnO3−δ, LSM) [50,64], Fe-based cathodes (e.g., SrFeO3−δ) [65,66], Ni-based cathodes (e.g., LaNiO3−δ), Co-based cathodes (e.g., (La,Sr)CoO3−δ) [36,39], and mixed transition metal-based cathodes (e.g., La(Ni,Fe)O3−δ and Sr(Co,Fe)O3−δ) [67,68]. Among them, cubic perovskite SrCoO3−δ-based cathodes inherited excellent combined ionic and electronic conductivity, yet they still suffer from insufficient structural stability at high temperature. To address this, various cation dopants (e.g., Ni, Cu, Zn, Cr, Fe, Al, Ga, In, Ce, Ti, Zr, Sn, V, Nb, etc.) are introduced to sustain cubic phase stability [52]. Recently developed high-performance Co-based and Fe-based cathode materials, such as SrCo0.9Nb0.1O3−δ (SCN), SrFe0.9Ti0.1O3−δ (SFT), and SrNb0.1Co0.9−yFeyO3−δ (SNCF) [69,70,71,72], perform particularly well in the low-temperature region (400–600 °C). Taking Co-based SrSc0.2Co0.8O3−δ (SSC) and SrNb0.1Co0.9O3−δ (SNC) as examples, they combine high conductivity with high oxygen vacancy concentration [73]. However, an enriched cobalt component leads to a larger thermal expansion coefficient (TEC), such as ~19.1 × 10−6 K−1 for SNC [74] and ~24 × 10−6 K−1 for SSC [75], posing an increased risk of cathode/electrolyte delamination during thermal cycling [76]. Additionally, CO2 sensitivity further exacerbates performance degradation.
To address this, Shao et al. systematically investigated the influence of the Co/Fe ratio at the B-site in SrNb0.1Co0.9−yFeyO3−δ (y = 0–0.9) in relation to the CO2 tolerance mechanism [64]. Their results indicate that the increase in the iron content significantly reduces the material’s sensitivity to CO2, which can be certified by the smaller increase in the area-specific resistance (ASR) after CO2 exposure. This phenomenon is attributed to the higher bond strength of Fe-O compared to Co-O, where the resulting increase in average metal–oxygen bond energy (ABE) effectively suppresses the carbonation reaction. Therefore, ABE can serve as a key descriptor for evaluating CO2 tolerance. Further studies confirmed that the substitution of the B-site of SrCoO3−δ with high-valence cations (e.g., Nb5+, Mo6+, Ti4+) can significantly enhance the perovskite’s CO2 tolerance [77,78,79]. This is verified by Zhu’s work, in which more negative ABE values denoted better CO2 tolerance, shown by the BSNM, LSCF, and BSCF (of which the ABEs were −414, −316, and 274 k mol−1, respectively), suggesting that BSNM has the best CO2 resistance (Figure 4a,b) [80]. In summary, the high ABE value induced by high-valence B-site cations is the fundamental factor in the improvement in CO2 tolerance.
Furthermore, Tejuca et al. investigated the extent of CO2 adsorption and the heat of adsorption on LaBO3 (B = Cr, Fe, Co) oxides over a wide temperature range [81]. They found that the chemical nature of the B-site cation significantly influences the CO2 coverage and its variation with the adsorption temperature. The results demonstrate that CO2 coverage follows the order LaCrO3 > LaFeO3 > LaCoO3. Consequently, the CO2 adsorption capacity is strongly dependent on the identity of the B-site cation.

3. Correlation Analysis Between SOFC Operating Parameters and Cathode CO2 Tolerance

Although the CO2 tolerance of perovskite cathodes is primarily governed by the elemental composition at the A/B-sites, the operating temperature also serves as a critical variable significantly impacting the performance. As shown in Figure 5a, Zhao et al. [47] investigated the CO2 tolerance of the La0.6Sr0.4CoO3−δ (LSC) cathode at different temperatures. The adsorption behavior in the low-temperature range (550–650 °C) accords with the Temkin model, while at a higher temperature range (700–800 °C), the model conforms to the Freundlich model. Notably, under constant CO2 partial pressure (e.g., 1.85 kPa), the increase in polarization resistance rises exponentially with decreasing temperature: only a 0.01 Ω cm2 increase at 700 °C, rising to 0.21 Ω cm2 at 600 °C, and surging to 1.20 Ω cm2 at 550 °C. This temperature effect is further reflected in the significantly higher rate of change in the area-specific resistance (ASR) with increasing CO2 partial pressure observed in the low-temperature range compared to the high-temperature range. Zhang et al. [61] studied the CO2 tolerance of SSNC and BSCF cathodes in atmospheres containing 10 vol% CO2 at 500–700 °C, as shown in Figure 5b. After exposure to 10 vol% CO2 at 500 °C and 600 °C, the relative increase rate of ASR reached 20–45 times, whereas at 700 °C, the relative increase rate was only 4–9 times. Furthermore, after CO2 removal from the atmosphere, the electrochemical performance of cathodes tested at 700 °C fully recovered within 15 min, while no complete recovery was observed for cathodes tested at 500 °C and 600 °C.
On perovskite cathodes, O2 preferentially adsorbs at B-site/oxygen-vacancy ensembles, forming O2(ad) to O2−/O via charge transfer before incorporation as O2−. In contrast, CO2 chemisorbs more strongly at A-site-terminated BaO/SrO domains, producing surface carbonate-like ACO3 species. These pathways compete for surface vacancies and electrons: carbonate formation blocks O2 adsorption/incorporation sites and reduces the population of mobile oxygen intermediates, thereby depressing the surface-exchange kinetics. The balance is dynamic: at higher temperature (or lower P C O 2 ), transient carbonates desorb, and O2 readsorbs rapidly (recovery), whereas at lower temperature and/or higher P C O 2 , persistent Ba/Sr-carbonates accumulate (poisoning). The schematic in Figure 5c visualizes these steps and the competition between CO2 and O2 on Ba/Sr-containing perovskites.
Nomura et al. discovered that the peritectic oxide formed in Ba0.95Ca0.05Co1−xFexO3−δ that filled with plenty of lattice oxygen at high temperatures readily adsorbs CO2 [82], suggesting that lattice oxygen in perovskites may catalyze carbonate formation. Significantly, the relationship between co-adsorption and competitive adsorption of CO2 and O2 undergoes a fundamental transformation with temperature. Yan et al.’s research on the BSCF system revealed that above 400 °C, CO2 and O2 exhibit significant competitive adsorption. When CO2 adheres on the cathode surface and occupies active sites, it significantly suppresses the adsorption, dissociation of oxygen molecules, and a reduction in adsorbed oxygen atoms, thereby degrading cell performance. However, in the high-temperature range of 600–700 °C, oxygen takes over the active sites on the surface and thereby leads the key reactions [83]. Collectively, these findings demonstrate that operating temperature not only modulates CO2 adsorption intensity but also governs the competitive relationship between CO2 and O2. Furthermore, thermodynamic data at lower temperatures indicate that materials containing Ba or Co elements are more susceptible to CO2 adsorption within perovskite structures [84]. In summary, increasing the operating temperature effectively mitigates CO2 poisoning effects.
When evaluating the CO2 tolerance of cathodes, there are significant differences in the test conditions adopted (e.g., CO2 partial pressure, exposure time, operating temperature), which poses challenges for direct performance comparisons among different material systems. Zhao et al. suggested that the CO2-poisoning effect is strongly dependent on parameters during tests. In their research on the La0.6Sr0.4CoO3−δ (LSC) cathode, a sharp increase in polarization resistance occurred when the CO2 partial pressure increased from 1.85 kPa to 4.68 kPa at 550 °C. However, at 700 °C, the effect of the same partial pressure change was much less significant, and the performance was quickly restored after CO2 removal [67]. The study was conducted under multiple temperature gradients (550–800 °C) and a series of precisely controlled CO2 partial pressures (0.53, 1.85, 2.99, 4.68 kPa). The coupling effect of temperature and CO2 partial pressure and its profound influence on the toxicity mechanism were clearly revealed. Therefore, to achieve fair and effective comparison of research results, future work urgently needs to clarify and unify the following core test parameters as much as possible when reporting CO2 tolerance.

4. Strategies for Enhancing Cathode Resistance to CO2 Poisoning

4.1. High-Acidity Cation Doping Strategy (Strategy via High-Acidity Cation Doping)

As discussed in the previous sections, the CO2 resistance of perovskite-type cathodes links to both A-site and B-site cations. The key influencing factors include ABE and the nature of alkaline earth metal elements. Crucially, the acidity level of the dopant cations is a decisive factor for CO2 tolerance (cation acidity order: Nb5+ > Co4+ > Co3+ > Fe3+ > Sc3+ > Co2+ > Fe2+ > Sr2+ > Ba2+). Introducing more acidic cations clearly enhances theCO2 resistance of perovskites.
Chen et al. [85] developed a calcium-doped double perovskite material, PrBa0.8Ca0.2Co2O5+δ (PBCC), which exhibits exceptional ORR activity and CO2 tolerance as a cathode. Almar et al. [86] demonstrated improved CO2 tolerance in BSCF through minor Y doping at the B-site. Focused ion beam (FIB) and scanning electron microscopy (SEM) analysis of the cell’s internal microstructure revealed that (Ba0.5Sr0.5)(Co0.8Fe0.2)0.9Y0.1O3−δ (BSCFY) exhibits significantly enhanced CO2 adsorption resistance in air at 700 °C. This improvement is primarily attributed to the higher acidity of Y3+ compared to A-site cations and the high concentration of Y dopants. Similarly, Huang et al. [87]. reported that Nb doping at the B-site enhances CO2 tolerance in BSCF. Stability measurements and XRD analysis showed that increasing the Nb doping content in Ba0.5Sr0.5Co0.8−xFe0.2NbxO3−δ (BSCFNbx) reduces BaCO3 peak intensity, indicating superior resistance to CO2 corrosion, likely due to the high acidity of Nb dopants. Kim et al. [88]. found that La3+ doping stabilizes the cubic perovskite structure in BSCF. Calculations based on the Goldschmidt tolerance factor indicate that substituting La3+ into the A-site of BSCF (forming La0.5Ba0.25Sr0.25Co0.8Fe0.2O3−δ, LBSCF) results in smaller cubic symmetry distortions, enhancing the structural stability and CO2 tolerance, while maintaining excellent phase stability under low p(O2) and CO2-containing atmospheres. Bi et al. [89] investigated Ti substitution for partial Co/Fe in BSCF as a cathode for proton-conducting SOFCs. The strong Ti-O bond reduces CO2 reactivity in Ti-doped BSCF, improving its chemical stability in pure CO2 at 600 °C. Li et al. [90] reported that partial substitution of Ba with Ca enhances CO2 tolerance in the proton-conducting Ba0.95Ca0.05Co0.4Fe0.4Zr0.1Y0.1O3−δ (BCaCFZY) cathode. Zhang et al. [91] documented that co-doping enhances ORR activity and CO2 resistance. The presence of 7.5 mol% Sc3+ promotes high oxygen valence states in Fe and oxygen vacancy formation in SrSc0.075Ta0.025Fe0.9O3−δ (SSTF75), enhancing ORR activity, while 2.5 mol% Ta5+ strengthens CO2 resistance due to its high valence state and acidity. Su et al. [92] reported that Ca-doped LaFeO3 cathodes exhibit good CO2 tolerance and low thermal expansion coefficients. DFT calculations revealed the CO2 adsorption mechanism on the surface, providing a strategy for designing Ca-doped lanthanum-based cathodes. In summary, doping strategies effectively enhance CO2 tolerance. Selected dopants should possess high acidity, high valence state, and high ABE characteristics; elements such as Ti4+, Zr4+, Nb5+, Sb5+, Ta5+, W6+, and Mo6+ show promise in mitigating CO2 poisoning in perovskites [93].
Taken together (Table 1), the available evidence suggests that higher valence, more Lewis-acidic B-site dopants, and/or stronger M-O bonds (e.g., Y3+/Nb5+/Ti4+/Ta5+) can improve apparent CO2 tolerance. That said, comparisons are complicated by the varying test conditions (1–10% CO2, 600–700 °C, 10 h) and by co-evolving factors such as tolerance-factor geometry, A-site basicity/Ba segregation, humidity, and conductivity. For example, Y-doped BSCF often shows smaller surface-exchange decay and less carbonate, but part of this advantage likely reflects phase stabilization; La on the A-site tends to stabilize the cubic phase (a geometric rather than acidity effect). Nb/Ti may weaken BaCO3 formation or improve chemical robustness yet can depress conductivity/ORR. Partial Ba to Ca substitution appears to suppress Ba segregation, especially under humid/protonic conditions. In Sc/Ta ferrites, CO2-induced ASR rises can be partly reversible, consistent with reversible surface coverage rather than persistent carbonate. Accordingly, “acidity-driven” improvements should be interpreted under matched conditions, with co-reporting of activity and stability to avoid over-attribution.

4.2. Composite SOFC Cathodes with Ceria-Based Electrolytes (SDC/GDC)

Incorporating electrolyte materials such as Sm0.2Ce0.8O1.9 (SDC) or Gd0.1Ce0.9O1.95 (GDC) into SOFC cathodes to form composite cathodes has been demonstrated as an effective approach to significantly enhance CO2 tolerance. The core mechanism involves the reduced relative content and surface exposure of alkaline earth metal elements (e.g., Sr and Ba) within the cathode phase after compositing with SDC/GDC. Simultaneously, SDC/GDC provides more stable and highly active ORR sites while forming physical barriers that restrict direct contact pathways between CO2 molecules and susceptible components in the cathode matrix at the microstructural level. This synergy improves the stability of the composite cathode in CO2-containing atmospheres. As shown in Figure 6a, Xue et al. proposed a CO2-tolerant stable oxygen-permeable membrane (60Ce0.9Gd0.1O2−δ-40Ba0.5Sr0.5Co0.8Fe0.2O3−δ) that exhibits higher oxygen permeability alongside superior stability in CO2-containing atmospheres compared to pure BSCF [93]. Li et al. found that mechanically incorporating samarium-doped ceria (SDC) into SrCo0.85Ta0.15O3−δ (SCT15) cathodes yielded composites with significantly lower sensitivity to 10% CO2 than pure SCT15 cathodes (Figure 6b) [94]. Thermogravimetric analysis, energy-dispersive spectroscopy, and conductivity data indicate that SDC nanoparticles improve triple-phase boundaries while simultaneously serving as protective layers. As shown in Figure 6c, Gu et al. studied SrCo0.8Nb0.1Ta0.1O3−δ (SCNT)-Ce0.9Gd0.1O2−δ (GDC) mechanically mixed composite cathodes, which exhibit enhanced CO2 resistance [95,96]. This system demonstrates superior ORR activity and lower thermal expansion coefficients compared to BSCF, with GDC blending reducing CO2-induced resistance sensitivity. Yang et al. [97] developed a cobalt-free Ba0.95La0.05FeO3−δ (BLF) cathode where the surface coating with GDC nanoparticles via mechanical mixing enhances O2 selectivity due to GDC’s high oxygen conductivity. This suppresses CO2 surface exchange, providing an effective approach to improve cathode performance and mitigate CO2 poisoning in air.
Adding SDC/GDC appears to improve CO2 tolerance (Table 2), plausibly through the chemical shielding of Sr/Ba sites and by maintaining wider and more TPB that help sustain ORR in CO2. The magnitude of the benefit seems sensitive to the architecture and loading: conformal coatings or percolating networks tend to perform better than simple mixes, whereas excessive ceria can dilute electronic pathways and increase ASR in clean air. Across both Co-containing and Co-free hosts, well-designed ceria phases often suppress ASR growth under CO2, albeit with a potential trade-off in peak ORR activity. Therefore, these observations point toward moderate ceria fractions implemented in optimized architectures as a pragmatic target.

4.3. Cathode Fabrication via Infiltration Strategy

The infiltration strategy has been validated as an effective approach for enhancing CO2 tolerance in SOFC cathodes. Qiu et al. [38] employed a solution infiltration technique to fabricate a novel core–shell structure for protecting BSCF. This design utilized LSM, known for its high catalytic activity and exceptional CO2 resistance, as a dense shell to encapsulate porous BSCF cathodes with high oxygen conductivity. This configuration effectively prevented a direct reaction between BSCF and CO2 while preserving optimal ORR activity and minimal polarization resistance (Rp) variation. Li et al. [79] demonstrated that infiltrating La2NiO4+δ (LNO) into a PrBa0.5Sr0.5Co1.5Fe0.5O5+δ (PBSCF) cathode significantly enhanced CO2 tolerance and ORR activity. The composite provided additional active sites for O2 adsorption and achieved near-complete performance recovery after CO2 removal. This improvement was attributed to LNO’s high oxygen exchange coefficient (k*), which facilitated dynamic oxygen surface exchange and localized CO2 adsorption on the LNO coating. Similarly, La0.94Ni0.6Fe0.4O3−δ (LNF94) [98] infiltration into La0.6Sr0.4Co0.2Fe0.8O3−δ (LSCF) cathodes suppressed Sr segregation and improved tolerance to CO2 and chromium poisoning. Pei et al. [99] reported that barium cobaltate (BCO)-coated LSCF cathodes exhibited enhanced ORR kinetics and CO2 resistance. Raman spectroscopy and DFT calculations confirmed that the BCO coating weakened CO2 adsorption and inhibited Sr surface segregation, thereby promoting ORR activity and mitigating CO2-induced degradation.
Taken together (Table 3), infiltration and coating appear to improve CO2 tolerance by forming thin conformal shells that may mask Sr/Ba-rich surfaces and potentially shift ORR toward faster surface-exchange sites. The benefit seems to depend on the loading and architecture. On LSCF, Ni-Fe perovskite (LNF94) at moderate loading has been reported to form a near-continuous film that lowers ASR and apparent activation energy, whereas heavier loading can lead to agglomeration, porosity loss, and an ASR rebound. In addition, on LSCF, BCO coatings have been observed to weaken CO2 adsorption, suppress the ~1066 cm−1 carbonate signal, and limit Sr segregation, thereby helping to stabilize ORR kinetics in CO2. For layered PBSCF, a La2NiO4+δ shell has been shown to reduce ASR and enable near-complete performance recovery after CO2 removal, in contrast to uncoated cathodes under similar conditions. Therefore, thin continuous infiltrates or percolating networks at moderate loading tend to be more effective than simply adding more material.

4.4. CO2-Induced In Situ Self-Reconstruction of the Cathode Surface

Enhancing cathode CO2 tolerance can also be achieved through surface in situ reconstruction, beyond conventional methods. As shown in Figure 7a, Zhou et al. [100] developed a novel cathode material, SrY0.05W0.05Co0.9O3−δ, and established an optimization strategy utilizing CO2-induced surface in situ reconstruction. Upon switching the treatment atmosphere from CO2 to air, nano-sized SrCO3 particles and oxide nanoparticles were generated. Both the SrCO3 and oxide nanoparticles were found to enhance ORR kinetics at the surface. Consequently, the ASR of a symmetric cell employing the SrY0.05W0.05Co0.05Co0.9O3−δ electrode decreased from 0.085 to 0.054 Ω cm2 at 600 °C.
Similarly, Shao et al. [101] proposed a SrSc0.025Nb0.075Fe0.9O3−δ (SSNF) perovskite, where CO2 exposure induced a beneficial transformation on the cathode surface (Figure 7b). The typically detrimental effect of carbonate formation was converted into a positive influence on the ORR activity for this perovskite. After treatment with high-concentration CO2 (10 vol%), a SrCO3 film formed on the perovskite surface, effectively shielding the underlying SSNF from further CO2 corrosion. This resulted in favorable and stable ORR activity even under high CO2 conditions (10 vol% in air). Furthermore, upon CO2 removal from the air, the surface carbonate film readily decomposed, leading to a significant improvement in ORR activity at intermediate temperatures.
Zhang et al. [40] discovered that CO2 treatment of a BaFeO3−δ (BF) cathode induced the in situ generation of barium carbonate, which reconstructed the cathode’s internal architecture, ultimately forming a high-performance core-shell structured cathode. This structure comprises a BaCO3 shell and a dual-phase core (containing BaFe2O4 and BaFeO3 phases), synergistically facilitating charge transfer, surface oxygen binding, and durability. Crucially, carbonate species can induce either surface activation or passivation of ORR on MIEC perovskites [72], depending on the treatment method, carbonate type, and loading ratio. The Ba(Co0.7Fe0.3)0.8Zr0.1Sc0.1O3−δ (BCFZS) cathode shows Zr/Sc co-doping tailors CO2 adsorption at Ba sites, causing beneficial nano-BaCO3 surface decoration instead of poisoning [102]. Simultaneously, in situ La/Co/Mg interdiffusion enhances the interface and oxygen-ion conduction.
Controlled CO2 exposure may transiently activate perovskite cathodes when it forms thin decomposable nano-carbonates that appear to protect or refresh the surface (e.g., SYWCo, SSNF, BF, BCFZS), whereas persistent coarser Ba/Sr carbonates tend to block transport and thus act as poisons. The balance seems to reflect a competition between A-site basicity (carbonate thermodynamics) and B-site acidity/oxygen-exchange kinetics (k*). Under 600–650 °C and 5–10 vol% CO2, brief treatments can favor the “activating” route, while lower temperatures, higher CO2, or dry feeds are more likely to promote passivation. Accordingly, strategies that seek to lower effective surface basicity while maintaining high k* may help bias reconstruction toward reversible nano-carbonates rather than bulk phases.
Beyond being recent, this review highlights CO2-induced in situ self-reconstruction as a distinct designable line of work by synthesizing scattered case reports into a tentative process window and material-selection notes. We also provide a side-by-side comparison of mainstream strategies (high-acidity/high-valence doping, composite cathodes, surface infiltration/coatings, and CO2-driven self-reconstruction) with operation complexity, advantages, and limitations to support practical choice (Table 4). While not exhaustive, we hope this framing offers a concise actionable overview for readers considering CO2 tolerance design.

5. Characterization

5.1. Operando Characterizations

Recent operando investigations directly connect CO2-induced surface chemistry with performance degradation in SOFC-relevant cathodes. Pagliari et al. [103] used operando NAP-XPS and synchrotron XRD to examine state-of-the-art cathodes at 700 °C (21% O2/79% CO2, up to 20 bar) and found SrCO3 on LSCF/LSC but not on LSM, with the performance largely reversible over 350 h under load (0.85 V). Opitz et al. [104] performed operando NAP-XPS on perovskite model electrodes during CO2 electrolysis (under cathodic polarization) and observed a high-binding-energy carbon species (~290 eV) consistent with a carbonate-type intermediate, requiring oxygen vacancies and electrons; graphitic carbon formed at strong bias could be removed upon depolarization. Yu et al. [105] conducted operando AP-XPS on ceria cells (~600 °C, 0.5 Torr CO/CO2, +2.0 V) and directly identified surface CO32− intermediates whose coverage increased under bias with concurrent Ce3+/Ce4+ changes, supporting a mechanism of pre-coordination to carbonate followed by electron-transfer reduction. Matras et al. [106] used (in situ/operando-style) XRD-CT on working BSCF membranes with CO2-containing feeds and reported a negative impact on product yields but no detectable carbonate phases; instead, degradation involved BSCF phase collapse and interphases including Ba6Co4O12 and BaWO4 (reaction-zone interfaces). Therefore, these operando results indicate that CO2 poisoning is governed by A-site-derived carbonate formation on Sr/Ba-containing perovskites and load-dependent surface redox; conversely, LSM-type or ceria-based surfaces show distinct responses. These insights suggest practical levers for material selection (limiting A-site basicity), bias/temperature windows, and surface-stabilization strategies.

5.2. DFT Calculations

Strengthening the role of theoretical guidance in the assessment and prediction of the chemical stability of cathode materials is of great importance. In addition to empirical stability assessment through exposure testing, computational methods based on density functional theory (DFT) provide profound insights into the intrinsic chemical stability of cathode materials from a fundamental perspective, offering a new paradigm beyond the traditional trial-and-error experimental approach. Dai et al. [107] demonstrated this by calculating the Gibbs free energy of the interaction between CO2 and Ca-doped LaMnO3 (LCaM) surfaces. Their DFT calculations incorporated vibrational entropy, and the results indicated that the adsorption of CO2 on the LCaM was thermodynamically unfavorable throughout the entire operating temperature range. This theoretical prediction of high stability was subsequently confirmed experimentally: after long-term exposure to a 600 °C CO2 atmosphere, the LCaM phase remained unchanged, and the fuel cell with an LCaM cathode exhibited highly stable performance. The strong agreement between computational predictions and experimental results shows that DFT-based energetics can serve as a powerful predictive descriptor for screening and designing CO2-tolerant cathode materials, thereby establishing a more rational foundation for material selection. DFT calculations provide key theoretical support for understanding the adsorption behavior of CO2 on the cathode surface. Studies have shown that the adsorption energy of CO2 on the surface of alkaline earth metal oxides (AEMO) follows the trend of BaO > SrO > CaO > MgO, which aligns with DFT-computed adsorption energies. DFT simulations further reveal that the adsorption energy of CO2 on AEMO surfaces is closely related to their basicity. Sites with stronger basicity (e.g., Ba2+) have a lower CO2 adsorption energy barrier and are thus more likely to form carbonates. During the chemical adsorption process, CO2 reacts with basic elements on the cathode surface (such as Ba2+, Sr2+) to form carbonates (such as SrCO3, BaCO3). DFT calculations indicate that the driving force of this process is closely related to the electronegativity of alkaline earth metal ions and the formation energy of oxygen vacancies. For instance, BaO has a relatively low oxygen vacancy formation energy, suggesting that its lattice oxygen is more readily involved in reactions, thereby promoting carbonate formation. In addition, DFT simulations have elucidated the dependence of CO2 adsorption on surface oxygen vacancy concentration: materials with higher oxygen vacancy concentrations are more susceptible to CO2 chemical adsorption, which in turn hinders oxygen adsorption and diffusion. Combined with thermodynamic analyses such as Ellingham diagrams, DFT has also demonstrated that Ba2+ possesses a higher driving force for carbonate formation than Sr2+ and La3+. It further quantifies alkalinity descriptors of A-site elements, such as local charge density and the oxygen p-band center and has revealed that the oxygen p-band center at Ba sites is closer to the Fermi level, indicating higher surface reactivity and a stronger tendency to interact with CO2. For example, Tsvetkov et al.’s research on La1−xSrxMnO3 DFT (LSM) materials shows that with the increase in the Sr doping amount, the CO2 in surface adsorption can significantly reduce, leading to more serious carbonate formation and attenuation of the performance [18]. This result is consistent with the experimental data of programmed temperature desorption (TPD), verifying the effectiveness of DFT in predicting the CO2 tolerance of materials. DFT can also serve to evaluate “acidity descriptors” for B-site cations, such as the average metal–oxygen bond energy (ABE) and B-site electronegativity. Studies indicate that high-valence and highly electronegative B-site cations (e.g., Nb5+, Mo6+, Ti4+) can significantly increase the ABE, thereby inhibiting CO2 adsorption and carbonate formation. For example, the DFT calculation of the SrNb0.1Co0.9−yFeyO3−δ system shows that as the Fe content increases, the ABE of the material rises, and the CO2 adsorption energy increases, resulting in the CO2 tolerance being enhanced [108]. Furthermore, DFT simulations have revealed the relationship between CO2 coverage and the type of B-site ion: in LaBO3 (B = Cr, Fe, Co), CO2 coverage follows the order LaCrO3 > LaFeO3 > LaCoO3, which is closely related to the electronic structure of the B-site ion and the activity of surface oxygen [109].

6. Conclusions and Perspectives

SOFCs face the significant challenge of cathode poisoning by atmospheric CO2 in real-life applications. CO2 adsorbs onto perovskite cathode surfaces, particularly reacting with alkaline earth metals such as barium, which exhibits far higher reactivity than strontium or lanthanum to form carbonates. This process not only inhibits surface oxygen exchange and the oxygen reduction reaction but also creates persistent carbonates that resist decomposition even after CO2 removal, critically impacting the long-term material stability. The interaction between CO2 and O2 at the cathode surface involves dynamic co-adsorption and competitive adsorption, which varies with temperature. Under competitive adsorption conditions, CO2 occupies valuable active sites, hindering oxygen molecule adsorption.
Despite notable progress in the study of CO2 poisoning, developing effective solutions remains a major challenge, due in part to insufficient systematic characterization and theoretical guidance. To address this, we propose the following perspectives and potential strategies to guide more in-depth investigations.
(1)
Developing new low-alkali/alkali-free perovskite cathode materials. The thermodynamic driving force of carbonate formation can be fundamentally avoided by designing and synthesizing high-performance perovskite cathode materials with no alkaline earth metals or low alkaline earth metal content. From this aspect, future research can focus on perovskite structures with rare earth elements (La, Pr, Sm, etc.) as the dominant A-sites.
(2)
Systematic screening and optimization of high-acidity/high-valence B-site dopants. High-acidity dopants are the key factor determining CO2 tolerance, and systematic evaluation of the mechanism will promote an in-depth understanding. Establishing the structural–activity relationship between dopants, structural stability, and CO2 tolerance by combining theoretical calculations and high-throughput experiments can open up novel strategies for synthesizing CO2-resistent perovskite.
(3)
Construct material systems with high ABE values. The value of ABE can be well used to evaluate the priority of carbonation reactions and thus can reflect the tolerance of the perovskite system to CO2. Priority should be given to developing material systems with high ABE values. For instance, by introducing elements such as Fe, Ti, and Zr to strengthen the B-O bond, the chemical adsorption of CO2 and the formation of carbonates can be inhibited. It is also expected to synthesize materials with high CO2 resistance.
(4)
Optimization of composite cathode structures and interface engineering. At present, research on composite cathodes mainly focuses on verifying the effectiveness, while in-depth study is required for the understanding of the microstructure of the composite interface, the uniformity of phase distribution, and long-term stability. In the future, efforts should be made to optimize the composite ratio and preparation process and to study at the microscopic scale how the SDC/GDC electrolyte phase acts as a physical barrier and highly active sites to precisely regulate gas transport and surface reaction pathways, so as to maximize the synergistic effect.
(5)
In-depth development of surface modification and in situ reconstruction strategies. Future research should focus on systematically revealing the thermodynamic boundaries and dynamic mechanisms of the reconstruction process. By precisely controlling the atmosphere, temperature, and time, the negative impact of CO2 is explored to be transformed into a controllable and beneficial surface self-assembly and activation process, thereby developing smart cathode materials with adaptive and self-repairing functions.
(6)
Innovation of operational strategies and system integration. By regularly introducing CO2-free air or controllable reducing atmosphere into the cathode can potentially restore the cathode activity by promoting carbonate decomposition; therefore, periodic atmosphere purging, temperature modulation, controlled reducing gas treatment, and other methods can be adopted to promote the decomposition and activity recovery of surface carbonates, and the impact of these dynamic strategies on the overall battery life, energy consumption, and system complexity should be systematically evaluated.
(7)
Strengthen the guidance of theoretical calculation in the assessment and prediction of the chemical stability of cathode materials. In addition to empirical stability assessment through exposure testing, the computational method based on density functional theory (DFT) provides profound insights into the intrinsic chemical stability of cathode materials from a fundamental perspective and offers a new angle for the traditional experimental trial-and-error approach. Dai et al. [107] demonstrated this by calculating the Gibbs free energy of the interaction between CO2 and Ca-doped LaMnO3 (LCaM) surfaces. Their DFT calculations incorporated vibrational entropy, and the results indicated that the adsorption of CO2 on the LCaM was thermodynamically unfavorable throughout the entire operating temperature range. This theoretical prediction of high stability was eventually confirmed by experimental observations: after long-term exposure to 600 °C CO2 atmosphere, the LCaM phase remained unchanged, and the fuel cell using the LCaM cathode exhibited highly stable output. The good consistency between computational predictions and experimental results indicates that the DFT-based energetics provides a powerful predictive descriptor for screening and designing CO2-resistant cathode materials, thereby laying a more reasonable foundation for material selection beyond the traditional trial-and-error method.
(8)
Application of practical operando tools coupled with electrochemical readouts. For CO2 poisoning, near-ambient pressure X-ray photoelectron spectroscopy (NAP/AP-XPS) (C 1s for carbonate/bicarbonate, O 1s for lattice/OH/CO32−, Sr/Ba core levels for A-site segregation) under realistic P C O 2 - P H 2 O -T-bias can be paired with synchrotron or lab in-situ XRD to follow phase changes, while operando Raman/DRIFTS tracks ν3(CO32−) and OH features. Complementary online MS (CO2/CO/H2O) together with simultaneous EIS/ASR helps relate surface chemistry to kinetics; occasional 13CO2/18O switching can clarify intermediate lifetimes and exchange routes. In related multi-pollutant contexts, in situ/operando SERS with mapping has already shown promise: for example, Chen et al. [110] detected and located SrCrO4 forming on LSCF surfaces in real time under operating conditions, providing direct evidence of the poisoning pathway. Building on such demonstrations, carefully designed broad-spectrum anti-poisoning strategies (targeting CO2, Cr, H2O) can be evaluated with the same operando toolset. To keep results comparable, we suggest reporting basic test histories ( P C O 2 , P H 2 O , T, flow, bias/time, and recovery steps) and presenting simple operating “maps” indicating where carbonate or other surface species persist or decompose. This measured approach should gradually clarify load- and humidity-dependent behavior on Sr/Ba-containing perovskites versus LSM/ceria-type surfaces and provide practical guidance for material choices and operating windows.
Therefore, enhancing long-term cathode CO2 tolerance demands both continuous material optimization (e.g., developing high-ABE high-acidity/high-valence dopants to simultaneously boost ORR activity and poisoning resistance) and the comprehensive evaluation and development of effective operational strategies. This integrated approach represents a crucial research direction for the practical implementation of SOFC technology.

Funding

This research was funded by the National Natural Science Foundation of China (NSFC), grant number 22109101 and the Shandong Energy Group, grant number SNKJ2022A01-R26.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Adler, S.B. Factors governing oxygen reduction in solid oxide fuel cell cathodes. Chem. Rev. 2004, 104, 4791–4843. [Google Scholar] [CrossRef]
  2. Cimenti, M.; Hill, J.M. Direct Utilization of Liquid Fuels in SOFC for Portable Applications: Challenges for the Selection of Alternative Anodes. Energies 2009, 2, 377–410. [Google Scholar] [CrossRef]
  3. Chitgar, N.; Moghimi, M. Design and evaluation of a novel multi-generation system based on SOFC-GT for electricity, fresh water and hydrogen production. Energy 2020, 197, 117162. [Google Scholar] [CrossRef]
  4. Chen, J.M.P.; Ni, M. Economic analysis of a solid oxide fuel cell cogeneration/trigeneration system for hotels in Hong Kong. Energy Build. 2014, 75, 160–169. [Google Scholar] [CrossRef]
  5. Steele, B.C.H. Materials for IT-SOFC stacks 35 years: The inevitability of gradualness? Solid State Ion. 2000, 134, 3–20. [Google Scholar] [CrossRef]
  6. Yang, Z.; Guo, M.; Wang, N.; Ma, C.; Wang, J.; Han, M. A short review of cathode poisoning and corrosion in solid oxide fuel cell. Int. J. Hydrogen Energy 2017, 42, 24948–24959. [Google Scholar] [CrossRef]
  7. Ivers-Tiffée, E.; Weber, A.; Herbstritt, D. Materials and technologies for SOFC-components. J. Eur. Ceram. Soc. 2001, 21, 1805–1811. [Google Scholar] [CrossRef]
  8. Malzbender, J.; Steinbrech, R.W. Advanced measurement techniques to characterize thermo-mechanical aspects of solid oxide fuel cells. J. Power Sources 2007, 173, 60–67. [Google Scholar] [CrossRef]
  9. Noviyanti, A.R.; Malik, Y.T.; Zain, S.M.; Risdiana. Figure-of-Merit Compatibility of Solid Oxide Fuel Cells. Energy Technol. 2023, 11, 2201418. [Google Scholar] [CrossRef]
  10. Exner, J.; Nazarenus, T.; Kita, J.; Moos, R. Dense Y-doped ion conducting perovskite films of BaZrO3, BaSnO3, and BaCeO3 for SOFC applications produced by powder aerosol deposition at room temperature. Int. J. Hydrogen Energy 2020, 45, 10000–10016. [Google Scholar] [CrossRef]
  11. Litzelman, S.J.; Lemmon, J.P. The Promise and Challenges of Intermediate Temperature Fuel Cells. ECS Trans. 2015, 68, 39. [Google Scholar] [CrossRef]
  12. Iwanschitz, B.; Sfeir, J.; Mai, A.; Schütze, M. Degradation of SOFC Anodes upon Redox Cycling: A Comparison Between Ni/YSZ and Ni/CGO. J. Electrochem. Soc. 2010, 157, B269–B278. [Google Scholar] [CrossRef]
  13. Li, Y.H.; Gemmen, R.; Liu, X.B. Oxygen reduction and transportation mechanisms in solid oxide fuel cell cathodes. J. Power Sources 2010, 195, 3345–3358. [Google Scholar] [CrossRef]
  14. Li, W.; Sunarso, J.; Yang, Y.; Chen, Y.; Ge, C.; Wang, W.; Guo, Y.; Ran, R.; Zhou, W. Strategies for improving oxygen ionic conducting in perovskite oxides and their practical applications. Energy Rev. 2024, 3, 100085. [Google Scholar] [CrossRef]
  15. Wang, H.; Cheng, J.; Chen, H.; Li, X.; Liu, D.; Li, X.; Li, W.; Wei, G.; Shen, S.; Chi, B.; et al. Applications of perovskite oxides for oxygen evolution and oxygen reduction reactions in alkaline media. Energy Rev. 2025, 4, 100139. [Google Scholar] [CrossRef]
  16. Boldrin, P.; Ruiz-Trejo, E.; Mermelstein, J.; Menéndez, J.M.B.; Reina, T.R.; Brandon, N.P. Strategies for Carbon and Sulfur Tolerant Solid Oxide Fuel Cell Materials, Incorporating Lessons from Heterogeneous Catalysis. Chem. Rev. 2016, 116, 13633–13684. [Google Scholar] [CrossRef] [PubMed]
  17. Shao, Z.; Haile, S.M. A high-performance cathode for the next generation of solid-oxide fuel cells. Nature 2004, 431, 170–173. [Google Scholar] [CrossRef] [PubMed]
  18. Tsvetkov, N.; Lu, Q.Y.; Sun, L.X.; Crumlin, E.J.; Yildiz, B. Improved chemical and electrochemical stability of perovskite oxides with less reducible cations at the surface. Nat. Mater. 2016, 15, 1010–1016. [Google Scholar] [CrossRef]
  19. Choi, S.; Davenport, T.C.; Haile, S.M. Protonic ceramic electrochemical cells for hydrogen production and electricity generation: Exceptional reversibility, stability, and demonstrated faradaic efficiency. Energy Environ. Sci. 2019, 12, 206–215. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Chen, B.; Guan, D.; Xu, M.; Ran, R.; Ni, M.; Zhou, W.; O’Hayre, R.; Shao, Z. Thermal-expansion offset for high-performance fuel cell cathodes. Nature 2021, 591, 246–251. [Google Scholar] [CrossRef]
  21. Zhang, Y.; Liu, Z.; Li, J.; Lin, K.; Guan, D.; Song, Y.; Yang, G.; Zhou, W.; Ge, J.; Shao, M.; et al. Interfacial oxide wedging for mechanical-robust electrode in high-temperature ceramic cells. Nat. Commun. 2025, 16, 8715. [Google Scholar] [CrossRef] [PubMed]
  22. Yang, H.X.; Zhang, Y.; Liu, Z.P.; Hu, C.F.; Li, J.B.; Liao, H.L.; Shao, M.H.; Ni, M.; Chen, B.; Shao, Z.P.; et al. Hydration-induced stiffness enabling robust thermal cycling of high temperature fuel cells cathode. Nat. Commun. 2025, 16, 3154. [Google Scholar] [CrossRef]
  23. Li, J.; Zhang, Y.; Zhu, H.; Yang, H.; Liu, Z.; Lin, K.; Sun, H.; Tian, Y.; Shen, S.; Xie, H.; et al. Mechanically-Enhanced, Single-Phased, and Triple-Conducting Air Electrode for Robust Oxygen-Ion and Proton Conducting Ceramic Cells. Adv. Funct. Mater. 2025, 35, 2502771. [Google Scholar] [CrossRef]
  24. Zhao, Z.; Liu, L.; Zhang, X.M.; Wu, W.M.; Tu, B.F.; Cui, D.A.; Ou, D.R.; Cheng, M.J. High- and low- temperature behaviors of La0.6Sr0.4Co0.2Fe0.8O3- δ cathode operating under CO2/H2O-containing atmosphere. Int. J. Hydrogen Energy 2013, 38, 15361–15370. [Google Scholar] [CrossRef]
  25. Qiu, P.; Lin, H.; Peng, Z.; Qi, H.; Tu, B.; Li, N.; Zhu, T.; Lu, Y.; Yang, Y. Research progress on the poisoning behavior of air electrodes and corresponding protection strategies for protonic ceramic cells. Energy Rev. 2025, 4, 100150. [Google Scholar] [CrossRef]
  26. Tofan, C.; Klvana, D.; Kirchnerova, J. Decomposition of nitric oxide over perovskite oxide catalysts: Effect of CO2, H2O and CH4. Appl. Catal. B-Environ. Energy 2002, 36, 311–323. [Google Scholar] [CrossRef]
  27. Bucher, E.; Egger, A.; Caraman, G.B.; Sitte, W. Stability of the SOFC Cathode Material (Ba,Sr)(Co,Fe)O3−δ in CO2-Containing Atmospheres. J. Electrochem. Soc. 2008, 155, B1218. [Google Scholar] [CrossRef]
  28. Partovi, K.; Liang, F.; Ravkina, O.; Caro, J. High-Flux Oxygen-Transporting Membrane Pr0.6Sr0.4Co0.5Fe0.5O3−δ: CO2 Stability and Microstructure. ACS Appl. Mater. Interfaces 2014, 6, 10274–10282. [Google Scholar] [CrossRef]
  29. Athanasiou, C.; Drosakis, C.; Booto, G.K.; Elmasides, C. Economic Feasibility of Power/Heat Cogeneration by Biogas–Solid Oxide Fuel Cell (SOFC) Integrated Systems. Energies 2023, 16, 404. [Google Scholar] [CrossRef]
  30. Han, Z.; Yang, Y.; Kong, D. Surface-scale affinity and adsorption selectivity of alkaline earth metal oxides to H2O and CO2: Insight into SOFC anode modification. Appl. Surf. Sci. 2020, 503, 144333. [Google Scholar] [CrossRef]
  31. Feldhoff, A.; Arnold, M.; Martynczuk, J.; Gesing, T.M.; Wang, H. The sol-gel synthesis of perovskites by an EDTA/citrate complexing method involves nanoscale solid state reactions. Solid State Sci. 2008, 10, 689–701. [Google Scholar] [CrossRef]
  32. Feldhoff, A.; Martynczuk, J.; Wang, H. Advanced Ba0.5Sr0.5Zn0.2Fe0.8O3−δ perovskite-type ceramics as oxygen selective membranes: Evaluation of the synthetic process. Prog. Solid State Chem. 2007, 35, 339–353. [Google Scholar] [CrossRef]
  33. Zhou, W.; Ran, R.; Shao, Z. Progress in understanding and development of Ba0.5Sr0.5Co0.8Fe0.2O3−δ-based cathodes for intermediate-temperature solid-oxide fuel cells: A review. J. Power Sources 2009, 192, 231–246. [Google Scholar] [CrossRef]
  34. Wang, Z.G.; Dewangan, N.; Das, S.; Wai, M.H.; Kawi, S. High oxygen permeable and CO2-tolerant SrCoxFe0.9-xNb0.1O3-δ(x=0.1-0.8) perovskite membranes: Behavior and mechanism. Sep. Purif. Technol. 2018, 201, 30–40. [Google Scholar] [CrossRef]
  35. Shao, Z.; Yang, W.; Cong, Y.; Dong, H.; Tong, J.; Xiong, G. Investigation of the permeation behavior and stability of a Ba0.5Sr0.5Co0.8Fe0.2O3−δ oxygen membrane. J. Membr. Sci. 2000, 172, 177–188. [Google Scholar] [CrossRef]
  36. Shao, Z.P.; Xiong, G.X.; Tong, J.H.; Dong, H.; Yang, W.S. Ba effect in doped Sr(Co0.8Fe0.2)O3-δ on the phase structure and oxygen permeation properties of the dense ceramic membranes. Sep. Purif. Technol. 2001, 25, 419–429. [Google Scholar] [CrossRef]
  37. Yan, A.Y.; Maragou, V.; Arico, A.; Cheng, M.; Tsiakaras, P. Investigation of a Ba0.5Sr0.5Co0.8Fe0.2O3- δ based cathode SOFC II.: The effect of CO2 on the chemical stability. Appl. Catal. B-Environ. 2007, 76, 320–327. [Google Scholar] [CrossRef]
  38. Qiu, P.; Wang, A.; Li, J.; Li, Z.B.; Jia, L.C.; Chi, B.; Pu, J.; Li, J. Promoted CO2-poisoning resistance of La0.8Sr0.2MnO3-δ-coated Ba0.5Sr0.5Co0.8Fe0.2O3- δ cathode for intermediate temperature solid oxide fuel cells. J. Power Sources 2016, 327, 408–413. [Google Scholar] [CrossRef]
  39. Yan, A.; Cheng, M.; Dong, Y.L.; Yang, W.S.; Maragou, V.; Song, S.Q.; Tsiakaras, P. Investigation of a Ba0.5Sr0.5Co0.8Fe0.2O3-δ based cathode IT-SOFC: I. The effect of CO2 on the cell performance. Appl. Catal. B Environ. 2006, 66, 64–71. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Li, J.; Xie, H.; Liu, Z.; Shen, S.; Teng, Y.; Guan, D.; Zhai, S.; Song, Y.; Zhou, W.; et al. CO2-induced reconstruction for ORR-enhanced solid oxide fuel cell cathode. Chem. Eng. J. 2023, 462, 142216. [Google Scholar] [CrossRef]
  41. Tan, X.Y.; Liu, N.; Meng, B.; Sunarso, J.; Zhang, K.; Liu, S.M. Oxygen permeation behavior of La0.6Sr0.4Co0.8Fe0.2O3 hollow fibre membranes with highly concentrated CO2 exposure. J. Membr. Sci. 2012, 389, 216–222. [Google Scholar] [CrossRef]
  42. Arratibel Plazaola, A.; Cruellas Labella, A.; Liu, Y.; Badiola Porras, N.; Pacheco Tanaka, D.A.; Sint Annaland, M.V.; Gallucci, F. Mixed Ionic-Electronic Conducting Membranes (MIEC) for Their Application in Membrane Reactors: A Review. Processes 2019, 7, 128. [Google Scholar] [CrossRef]
  43. Jiang, S.P. A comparison of O2 reduction reactions on porous (La,Sr)MnO3 and (La,Sr)(Co,Fe)O3 electrodes. Solid State Ion. 2002, 146, 1–22. [Google Scholar] [CrossRef]
  44. Xia, C.R.; Rauch, W.; Chen, F.L.; Liu, M.L. Sm0.5Sr0.5CoO3 cathodes for low-temperature SOFCs. Solid State Ion. 2002, 149, 11–19. [Google Scholar] [CrossRef]
  45. Choi, S.; Yoo, S.; Kim, J.; Park, S.; Jun, A.; Sengodan, S.; Kim, J.; Shin, J.; Jeong, H.Y.; Choi, Y.; et al. Highly efficient and robust cathode materials for low-temperature solid oxide fuel cells: PrBa0.5Sr0.5Co2-xFexO5+δ. Sci. Rep. 2013, 3, srep02426. [Google Scholar] [CrossRef]
  46. Chen, D.J.; Shao, Z.P. Surface exchange and bulk diffusion properties of Ba0.5Sr0.5Co0.8Fe0.2O3- δ mixed conductor. Int. J. Hydrogen Energy 2011, 36, 6948–6956. [Google Scholar] [CrossRef]
  47. Zhao, Z.; Liu, L.; Zhang, X.M.; Wu, W.M.; Tu, B.F.; Ou, D.R.; Cheng, M.J. A comparison on effects of CO2 on La0.8Sr0.2MnO3+δ and La0.6Sr0.4CoO3-δ cathodes. J. Power Sources 2013, 222, 542–553. [Google Scholar] [CrossRef]
  48. Zhang, H.; Cong, Y.; Yang, W. Effect of CO2 Treatment on the Performance of Sm0.5Sr0.5CoO3-δ Cathode Electrocatalyst. Chin. J. Catal. 2008, 29, 7–9. [Google Scholar] [CrossRef]
  49. Kharton, V.V.; Yaremchenko, A.A.; Kovalevsky, A.V.; Viskup, A.P.; Naumovich, E.N.; Kerko, P.F. Perovskite-type oxides for high-temperature oxygen separation membranes. J. Membr. Sci. 1999, 163, 307–317. [Google Scholar] [CrossRef]
  50. Dong, F.; Chen, D.; Ran, R.; Park, H.; Kwak, C.; Shao, Z. A comparative study of Sm0.5Sr0.5MO3−δ (M = Co and Mn) as oxygen reduction electrodes for solid oxide fuel cells. Int. J. Hydrogen Energy 2012, 37, 4377–4387. [Google Scholar] [CrossRef]
  51. Kim, J.H.; Baek, S.W.; Lee, C.; Park, K.; Bae, J. Performance analysis of cobalt-based cathode materials for solid oxide fuel cell. Solid State Ion. 2008, 179, 1490–1496. [Google Scholar] [CrossRef]
  52. Nagai, T.; Ito, W.; Sakon, T. Relationship between cation substitution and stability of perovskite structure in SrCoO3–δ-based mixed conductors. Solid State Ion. 2007, 177, 3433–3444. [Google Scholar] [CrossRef]
  53. Yi, J.; Feng, S.; Zuo, Y.; Liu, W.; Chen, C. Oxygen Permeability and Stability of Sr0.95Co0.8Fe0.2O3-δ in a CO2- and H2O-Containing Atmosphere. ChemInform 2006, 37, e21. [Google Scholar] [CrossRef]
  54. Khromushin, I.V.; Aksenova, T.I.; Zhotabaev, Z.R. Mechanism of gas-solid exchange processes for some perovskites. Solid State Ion. 2003, 162, 37–40. [Google Scholar] [CrossRef]
  55. Zakowsky, N.; Williamson, S.; Irvine, J.T.S. Elaboration of CO2 tolerance limits of BaCe0.9Y0.1O3-δ electrolytes for fuel cells and other applications. Solid State Ion. 2005, 176, 3019–3026. [Google Scholar] [CrossRef]
  56. Kaus, I.; Wiik, K.; Krogh, B.; Dahle, M.; Hofstad, K.H.; Aasland, S. Stability of SrFeO3-based materials in H2O/CO2-containing atmospheres at high temperatures and pressures. J. Am. Ceram. Soc. 2007, 90, 2226–2230. [Google Scholar] [CrossRef]
  57. Wang, H.; Tablet, C.; Yang, W.; Caro, J. In situ high temperature X-ray diffraction studies of mixed ionic and electronic conducting perovskite-type membranes. Mater. Lett. 2005, 59, 3750–3755. [Google Scholar] [CrossRef]
  58. Meng, B.; Wang, Z.G.; Liu, Y.Y.; Tan, X.Y.; da Costa, J.C.D.; Liu, S.M. Preparation and oxygen permeation properties of SrCo0.9Nb0.1O3-δ hollow fibre membranes. Sep. Purif. Technol. 2011, 78, 175–180. [Google Scholar] [CrossRef]
  59. Meng, B.; Wang, Z.G.; Tan, X.Y.; Liu, S.M. SrCo0.9Sc0.1O3-δ perovskite hollow fibre membranes for air separation at intermediate temperatures. J. Eur. Ceram. Soc. 2009, 29, 2815–2822. [Google Scholar] [CrossRef]
  60. Wang, Z.G.; Kathiraser, Y.; Kawi, S. High performance oxygen permeable membranes with Nb-doped BaBi0.05Co0.95O3-δ perovskite oxides. J. Membr. Sci. 2013, 431, 180–186. [Google Scholar] [CrossRef]
  61. Zhang, Y.; Yang, G.M.; Chen, G.; Ran, R.; Zhou, W.; Shao, Z.P. Evaluation of the CO2 Poisoning Effect on a Highly Active Cathode SrSc0.175Nb0.025Co0.8O3-δ in the Oxygen Reduction Reaction. Acs Appl. Mater. Interfaces 2016, 8, 3003–3011. [Google Scholar] [CrossRef]
  62. Hu, B.X.; Mahapatra, M.K.; Keane, M.; Zhang, H.; Singh, P. Effect of CO2 on the stability of strontium doped lanthanum manganite cathode. J. Power Sources 2014, 268, 404–413. [Google Scholar] [CrossRef]
  63. Klande, T.; Efimov, K.; Cusenza, S.; Becker, K.-D.; Feldhoff, A. Effect of doping, microstructure, and CO2 on La2NiO4+δ-based oxygen-transporting materials. J. Solid State Chem. 2011, 184, 3310–3318. [Google Scholar] [CrossRef]
  64. Liang, F.L.; Chen, J.; Jiang, S.P.; Chi, B.; Pu, J.; Jian, L. High performance solid oxide fuel cells with electrocatalytically enhanced (La, Sr)MnO3 cathodes. Electrochem. Commun. 2009, 11, 1048–1051. [Google Scholar] [CrossRef]
  65. Huang, Y.Y.; Vohs, J.M.; Gorte, R.J. Fabrication of Sr-doped LaFeO3 YSZ composite cathodes. J. Electrochem. Soc. 2004, 151, A646–A651. [Google Scholar] [CrossRef]
  66. Orikasa, Y.; Nakao, T.; Oishi, M.; Ina, T.; Mineshige, A.; Amezawa, K.; Arai, H.; Ogumi, Z.; Uchimoto, Y. Local structural analysis for oxide ion transport in La0.6Sr0.4FeO3-δ cathodes. J. Mater. Chem. 2011, 21, 14013–14019. [Google Scholar] [CrossRef]
  67. Liu, H.Y.; Zhu, X.F.; Cong, Y.; Zhang, T.Y.; Yang, W.S. Remarkable dependence of electrochemical performance of SrCo0.8Fe0.2O3-δ on A-site nonstoichiometry. Phys. Chem. Chem. Phys. 2012, 14, 7234–7239. [Google Scholar] [CrossRef]
  68. Orui, H.; Watanabe, K.; Chiba, R.; Arakawa, M. Application of LaNi(Fe)O3 as SOFC Cathode. J. Electrochem. Soc. 2004, 151, A1412. [Google Scholar] [CrossRef]
  69. Baharuddin, N.A.; Muchtar, A.; Somalu, M.R.; Samat, A.A. Thermal Decomposition of Cobalt-free SrFe0.9Ti0.1O3-δ Cathode for Intermediate Temperature Solid Oxide Fuel Cell. Procedia Eng. 2016, 148, 72–77. [Google Scholar] [CrossRef]
  70. Jiang, S.; Sunarso, J.; Zhou, W.; Shen, J.; Ran, R.; Shao, Z. Cobalt-free SrNbxFe1−xO3−δ (x = 0.05, 0.1 and 0.2) perovskite cathodes for intermediate temperature solid oxide fuel cells. J. Power Sources 2015, 298, 209–216. [Google Scholar] [CrossRef]
  71. Chen, Y.B.; Qian, B.M.; Shao, Z.P. Tin and iron co-doping strategy for developing active and stable oxygen reduction catalysts from SrCoO3-δ for operating below 800 °C. J. Power Sources 2015, 294, 339–346. [Google Scholar] [CrossRef]
  72. Zhu, Y.L.; Sunarso, J.; Zhou, W.; Shao, Z.P. Probing CO2 reaction mechanisms and effects on the SrNb0.1Co0.9-xFexO3- δ cathodes for solid oxide fuel cells. Appl. Catal. B-Environ. 2015, 172, 52–57. [Google Scholar] [CrossRef]
  73. Zhou, W.; Shao, Z.P.; Ran, R.; Cai, R. Novel SrSc0.2Co0.8O3-δ as a cathode material for low temperature solid-oxide fuel cell. Electrochem. Commun. 2008, 10, 1647–1651. [Google Scholar] [CrossRef]
  74. Zhou, W.; Jin, W.Q.; Zhu, Z.H.; Shao, Z.P. Structural, electrical and electrochemical characterizations of SrNb0.1Co0.9O3-δ as a cathode of solid oxide fuel cells operating below 600 °C. Int. J. Hydrogen Energy 2010, 35, 1356–1366. [Google Scholar] [CrossRef]
  75. Fukunaga, H.; Koyama, M.; Takahashi, N.; Wen, C.; Yamada, K. Reaction model of dense Sm0.5Sr0.5CoO3 as SOFC cathode. Solid State Ion. 2000, 132, 279–285. [Google Scholar] [CrossRef]
  76. Zając, W.; Świerczek, K.; Molenda, J. Thermochemical compatibility between selected (La,Sr)(Co,Fe,Ni)O3 cathodes and rare earth doped ceria electrolytes. J. Power Sources 2007, 173, 675–680. [Google Scholar] [CrossRef]
  77. Wang, R.; Jin, F.; Ta, L.; He, T. SrCo1−xMoxO3−δ perovskites as cathode materials for LaGaO3-based intermediate-temperature solid oxide fuel cells. Solid State Ion. 2016, 288, 32–35. [Google Scholar] [CrossRef]
  78. Yang, G.M.; Su, C.; Chen, Y.B.; Dong, F.F.; Tade, M.O.; Shao, Z.P. Cobalt-free SrFe0.9Ti0.1O3-δ as a high-performance electrode material for oxygen reduction reaction on doped ceria electrolyte with favorable CO2 tolerance. J. Eur. Ceram. Soc. 2015, 35, 2531–2539. [Google Scholar] [CrossRef]
  79. Li, J.; Zhang, Q.; Qiu, P.; Jia, L.; Chi, B.; Pu, J.; Li, J. A CO2-tolerant La2NiO4+δ-coated PrBa0.5Sr0.5Co1.5Fe0.5O5+δ cathode for intermediate temperature solid oxide fuel cells. J. Power Sources 2017, 342, 623–628. [Google Scholar] [CrossRef]
  80. Zhu, Y.; Zhou, W.; Chen, Y.; Shao, Z. An Aurivillius Oxide Based Cathode with Excellent CO2 Tolerance for Intermediate-Temperature Solid Oxide Fuel Cells. Angew. Chem. Int. Ed. 2016, 55, 8988–8993. [Google Scholar] [CrossRef]
  81. Peña, M.A.; Fierro, J.L.G. Chemical Structures and Performance of Perovskite Oxides. Chem. Rev. 2001, 101, 1981–2018. [Google Scholar] [CrossRef]
  82. Nomura, K.; Ujihira, Y.; Hayakawa, T.; Takehira, K. CO2 absorption properties and characterization of perovskite oxides, (Ba,Ca) (Co,Fe) O3−δ. Appl. Catal. A Gen. 1996, 137, 25–36. [Google Scholar] [CrossRef]
  83. Yan, A.Y.; Bin, L.; Dong, Y.L.; Tian, Z.J.; Wang, D.Z.; Cheng, M.J. A temperature programmed desorption investigation on the interaction of Ba0.5Sr0.5Co0.8>Fe0.2O3- δ perovskite oxides with CO2 in the absence and presence of H2O and O2. Appl. Catal. B-Environ. 2008, 80, 24–31. [Google Scholar] [CrossRef]
  84. Scholten, M.J.; Schoonman, J.; van Miltenburg, J.C.; Oonk, H.A.J. Synthesis of strontium and barium cerate and their reaction with carbon dioxide. Solid State Ion. 1993, 61, 83–91. [Google Scholar] [CrossRef]
  85. Chen, Y.; Yoo, S.; Choi, Y.; Kim, J.H.; Ding, Y.; Pei, K.; Murphy, R.; Zhang, Y.; Zhao, B.; Zhang, W.; et al. A highly active, CO2-tolerant electrode for the oxygen reduction reaction. Energy Environ. Sci. 2018, 11, 2458–2466. [Google Scholar] [CrossRef]
  86. Almar, L.; Störmer, H.; Meffert, M.; Szász, J.; Wankmüller, F.; Gerthsen, D.; Ivers-Tiffée, E. Improved Phase Stability and CO2 Poisoning Robustness of Y-Doped Ba0.5Sr0.5Co0.8Fe0.2O3−δ SOFC Cathodes at Intermediate Temperatures. ACS Appl. Energy Mater. 2018, 1, 1316–1327. [Google Scholar] [CrossRef]
  87. Huang, Y.W.; Ding, J.W.; Xia, Y.P.; Miao, L.N.; Li, K.; Zhang, Q.P.; Liu, W. Ba0.5Sr0.5C0.8-xFe0.2NbxO3-δ (x ≤ 0.1) as cathode materials for intermediate temperature solid oxide fuel cells with an electron-blocking interlayer. Ceram. Int. 2020, 46, 10215–10223. [Google Scholar] [CrossRef]
  88. Kim, J.; Choi, S.; Jun, A.; Jeong, H.Y.; Shin, J.; Kim, G. Chemically Stable Perovskites as Cathode Materials for Solid Oxide Fuel Cells: La-Doped Ba0.5Sr0.5Co0.8Fe0.2O3−δ. ChemSusChem 2014, 7, 1669–1675. [Google Scholar] [CrossRef] [PubMed]
  89. Bi, L.; Fabbri, E.; Traversa, E. Novel Ba0.5Sr0.5(Co0.8Fe0.2)1−xTixO3−δ (x = 0, 0.05, and 0.1) cathode materials for proton-conducting solid oxide fuel cells. Solid State Ion. 2012, 214, 1–5. [Google Scholar] [CrossRef]
  90. Li, J.; Hou, J.; Lu, Y.; Wang, Q.; Xi, X.; Fan, Y.; Fu, X.-Z.; Luo, J.-L. Ca-containing Ba0·95Ca0·05Co0·4Fe0·4Zr0·1Y0·1O3-δ cathode with high CO2-poisoning tolerance for proton-conducting solid oxide fuel cells. J. Power Sources 2020, 453, 227909. [Google Scholar] [CrossRef]
  91. Zhang, Y.; Gao, X.; Sunarso, J.; Liu, B.; Zhou, W.; Ni, M.; Shao, Z. Significantly Improving the Durability of Single-Chamber Solid Oxide Fuel Cells: A Highly Active CO2-Resistant Perovskite Cathode. ACS Appl. Energy Mater. 2018, 1, 1337–1343. [Google Scholar] [CrossRef]
  92. Su, M.; Huan, D.; Hu, X.; Zhu, K.; Peng, R.; Xia, C. Understanding the favorable CO2 tolerance of Ca-doped LaFeO3 perovskite cathode for solid oxide fuel cells. J. Power Sources 2022, 521, 230907. [Google Scholar] [CrossRef]
  93. Xue, J.; Liao, Q.; Wei, Y.; Li, Z.; Wang, H. A CO2-tolerance oxygen permeable 60Ce0.9Gd0.1O2−δ–40Ba0.5Sr0.5Co0.8Fe0.2O3−δ dual phase membrane. J. Membr. Sci. 2013, 443, 124–130. [Google Scholar] [CrossRef]
  94. Li, M.; Zhou, W.; Zhu, Z. Highly CO2-Tolerant Cathode for Intermediate-Temperature Solid Oxide Fuel Cells: Samarium-Doped Ceria-Protected SrCo0.85Ta0.15O3−δ Hybrid. ACS Appl. Mater. Interfaces 2017, 9, 2326–2333. [Google Scholar] [CrossRef] [PubMed]
  95. Li, M.; Zhao, X.; Min, H.; Yuan, G.; Ding, X. Synergistically enhancing CO2-tolerance and oxygen reduction reaction activity of cobalt-free dual-phase cathode for solid oxide fuel cells. Int. J. Hydrogen Energy 2020, 45, 34058–34068. [Google Scholar] [CrossRef]
  96. Gu, B.; Sunarso, J.; Zhang, Y.; Song, Y.; Yang, G.; Zhou, W.; Shao, Z. A high performance composite cathode with enhanced CO2 resistance for low and intermediate-temperature solid oxide fuel cells. J. Power Sources 2018, 405, 124–131. [Google Scholar] [CrossRef]
  97. Yang, K.C.; Wang, Y.H.; Yang, Z.B.; Wang, Y.R.; Jin, C. Degradation of cobalt-free Ba0.95La0.05FeO3-δ cathode against CO2-H2O and Ce0.9Gd0.1O2-δ modification. Int. J. Hydrogen Energy 2020, 45, 34080–34088. [Google Scholar] [CrossRef]
  98. Zhao, H.; Li, W.; Wang, H.; Zhou, J.; Sun, X.; Wang, E.; Zhao, L.; Dong, B.; Wang, S. Surface modification of La0.6Sr0.4Co0.2Fe0.8O3 cathode by infiltrating A-site deficient non-strontium La0.94Ni0.6Fe0.4O3 perovskite for solid oxide fuel cells. Appl. Surf. Sci. 2022, 572, 151382. [Google Scholar] [CrossRef]
  99. Pei, K.; Zhou, Y.C.; Ding, Y.; Xu, K.; Zhang, H.; Yuan, W.; Sasaki, K.; Choi, Y.M.; Liu, M.L.; Chen, Y. An improved oxygen reduction reaction activity and CO2-tolerance of La0.6Sr0.4Co0.2Fe0.8O3-δ achieved by a surface modification with barium cobaltite coatings. J. Power Sources 2021, 514, 230573. [Google Scholar] [CrossRef]
  100. Liu, D.; Chen, W.; Zhou, C.; Fei, M.; Liang, F.; Gu, Y.; Xu, M.; Ran, R.; Zhou, W. CO2-induced in-situ surface reconfiguration of strontium cobaltite-based perovskite for accelerated oxygen reduction reaction. Appl. Surf. Sci. 2023, 629, 157452. [Google Scholar] [CrossRef]
  101. Gu, H.X.; Sunarso, J.; Yang, G.M.; Zhou, C.; Song, Y.F.; Zhang, Y.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z.P. Turning Detrimental Effect into Benefits: Enhanced Oxygen Reduction Reaction Activity of Cobalt-Free Perovskites at Intermediate Temperature via CO2-Induced Surface Activation. Acs Appl. Mater. Interfaces 2020, 12, 16417–16425. [Google Scholar] [CrossRef]
  102. Wang, H.; Zhu, W.; Gong, J.; Guan, M.; Xu, L.; Zhang, W.; Liu, X.; Yan, C.; Meng, J. Effect of octahedral occupancy of bimetal-doping and CO2-induced surface reconstruction on oxygen reduction reaction of cobalt-based perovskite oxides. Chem. Eng. J. 2024, 485, 149770. [Google Scholar] [CrossRef]
  103. Pagliari, M.; Marasi, M.; Donazzi, A. Electrocatalytic and operando characterization of state-of-the-art SOFC cathodes for applications at high CO2 concentration in novel clean power production cycles. Electrochim. Acta 2024, 501, 144779. [Google Scholar] [CrossRef]
  104. Opitz, A.K.; Nenning, A.; Rameshan, C.; Kubicek, M.; Götsch, T.; Blume, R.; Hävecker, M.; Knop-Gericke, A.; Rupprechter, G.; Klötzer, B.; et al. Surface Chemistry of Perovskite-Type Electrodes During High Temperature CO2 Electrolysis Investigated by Operando Photoelectron Spectroscopy. ACS Appl. Mater. Interfaces 2017, 9, 35847–35860. [Google Scholar] [CrossRef] [PubMed]
  105. Yu, Y.; Mao, B.H.; Geller, A.; Chang, R.; Gaskell, K.; Liu, Z.; Eichhorn, B.W. CO2 activation and carbonate intermediates: An operando AP-XPS study of CO2 electrolysis reactions on solid oxide electrochemical cells. Phys. Chem. Chem. Phys. 2014, 16, 11633–11639. [Google Scholar] [CrossRef]
  106. Matras, D.; Vamvakeros, A.; Jacques, S.D.M.; Middelkoop, V.; Vaughan, G.; Aran, M.A.; Cernik, R.J.; Beale, A.M. In situ X-ray diffraction computed tomography studies examining the thermal and chemical stabilities of working Ba0.5Sr0.5Co0.8Fe0.2O3-δ membranes during oxidative coupling of methane. Phys. Chem. Chem. Phys. 2020, 22, 18964–18975. [Google Scholar] [CrossRef]
  107. Dai, H.; Xu, X.; Liu, C.; Ma, C.; Zhang, Q.; Bi, L. Tailoring a LaMnO3 cathode for proton-conducting solid oxide fuel cells: Integration of high performance and excellent stability. J. Mater. Chem. A 2021, 9, 12553–12559. [Google Scholar] [CrossRef]
  108. Zhu, K.Y.; Liu, H.Y.; Li, X.N.; Li, Q.M.; Wang, J.H.; Zhu, X.F.; Yang, W.S. Oxygen evolution reaction over Fe site of BaZrxFe1-xO3-δ perovskite oxides. Electrochim. Acta 2017, 241, 433–439. [Google Scholar] [CrossRef]
  109. Zhang, S.; Han, N.; Tan, X. Density functional theory calculations of atomic, electronic and thermodynamic properties of cubic LaCoO3 and La1−xSrxCoO3 surfaces. RSC Adv. 2015, 5, 760–769. [Google Scholar] [CrossRef]
  110. Chen, Y.; Yoo, S.; Li, X.; Ding, D.; Pei, K.; Chen, D.; Ding, Y.; Zhao, B.; Murphy, R.; deGlee, B.; et al. An effective strategy to enhancing tolerance to contaminants poisoning of solid oxide fuel cell cathodes. Nano Energy 2018, 47, 474–480. [Google Scholar] [CrossRef]
Figure 3. (a). Desorption of CO2 from La1−xSrxMnO3−δ annealed in air. Redrawn from Ref. [54], (b). Scanning electron microscopy (SEM) images of the surface of (a) SSNC-D and (b) BSCF-D before and after CO2 treatment at 500, 600, and 700 °C. Reproduced with permission from Ref. [61]. (c). Schematic of the mechanisms of CO2 interaction with LSM and related electrochemical performance. A: as-fabricated LSM cathode; B: SrO incorporation into the LSM lattice; LSM activation in dry air. C: No more SrCO3 formation at the LSM surface due to absence of SrO; stable electrochemical performance; D: partial incorporation of SrO into the LSM lattice and formation SrCO3 at the LSM surface; limited activation and performance degradation, E: increased SrCO3 on the surface and subsequent performance degradation; an illustration from Ref. [62].
Figure 3. (a). Desorption of CO2 from La1−xSrxMnO3−δ annealed in air. Redrawn from Ref. [54], (b). Scanning electron microscopy (SEM) images of the surface of (a) SSNC-D and (b) BSCF-D before and after CO2 treatment at 500, 600, and 700 °C. Reproduced with permission from Ref. [61]. (c). Schematic of the mechanisms of CO2 interaction with LSM and related electrochemical performance. A: as-fabricated LSM cathode; B: SrO incorporation into the LSM lattice; LSM activation in dry air. C: No more SrCO3 formation at the LSM surface due to absence of SrO; stable electrochemical performance; D: partial incorporation of SrO into the LSM lattice and formation SrCO3 at the LSM surface; limited activation and performance degradation, E: increased SrCO3 on the surface and subsequent performance degradation; an illustration from Ref. [62].
Processes 13 03931 g003
Figure 4. (a). Impedance spectra of BSNM, LSCF, and BSCF cathodes at 700 °C in air and in air containing 10 vol % CO2. Redrawn from [80]. (b). Average metal-oxygen bond energies (ABEs) for the BSNM, LSCF, and BSCF cathodes. Reproduced with permission from Ref. [80].
Figure 4. (a). Impedance spectra of BSNM, LSCF, and BSCF cathodes at 700 °C in air and in air containing 10 vol % CO2. Redrawn from [80]. (b). Average metal-oxygen bond energies (ABEs) for the BSNM, LSCF, and BSCF cathodes. Reproduced with permission from Ref. [80].
Processes 13 03931 g004
Figure 5. (a). Electrochemical impedance spectra of the LSC cell at various CO2 partial pressures at 550 °C. Redrawn from Ref. [47]. (b). Relative changes in the ASRs of the SSNC and BSCF cathodes after introduction of CO2 to air at different temperatures. The “removal” indicates the ASRs of cathodes after removing the CO2 from air for 15 min. Reproduced with permission from Ref. [61]. (c). Schematic of dynamic co-adsorption and competition between CO2 and O2 on perovskite cathodes: O2 adsorbs/dissociates at B-site/vacancy ensembles and incorporates as O2−; CO2 chemisorbs to form surface carbonates, blocking oxygen-exchange sites and toggling between reversible (transient) and persistent (blocking) regimes.
Figure 5. (a). Electrochemical impedance spectra of the LSC cell at various CO2 partial pressures at 550 °C. Redrawn from Ref. [47]. (b). Relative changes in the ASRs of the SSNC and BSCF cathodes after introduction of CO2 to air at different temperatures. The “removal” indicates the ASRs of cathodes after removing the CO2 from air for 15 min. Reproduced with permission from Ref. [61]. (c). Schematic of dynamic co-adsorption and competition between CO2 and O2 on perovskite cathodes: O2 adsorbs/dissociates at B-site/vacancy ensembles and incorporates as O2−; CO2 chemisorbs to form surface carbonates, blocking oxygen-exchange sites and toggling between reversible (transient) and persistent (blocking) regimes.
Processes 13 03931 g005
Figure 6. (a). Diagrams of the novel concept for CO2-tolerance oxygen-permeable dual-phase membranes. Redrawn from Ref. [93]. (b). Relative ASR value changes with time for the studied cathodes when the flowing air is replaced by flowing 10% CO2-containing air at 550 °C. Reproduced with permission from Ref. [94]. (c). Electrochemical impedance spectra evolution during 100 h test at 550 °C in 1 vol % CO2-containing air for SCNT and SCNT-GDC cathode. Reproduced with permission from Ref. [96].
Figure 6. (a). Diagrams of the novel concept for CO2-tolerance oxygen-permeable dual-phase membranes. Redrawn from Ref. [93]. (b). Relative ASR value changes with time for the studied cathodes when the flowing air is replaced by flowing 10% CO2-containing air at 550 °C. Reproduced with permission from Ref. [94]. (c). Electrochemical impedance spectra evolution during 100 h test at 550 °C in 1 vol % CO2-containing air for SCNT and SCNT-GDC cathode. Reproduced with permission from Ref. [96].
Processes 13 03931 g006
Figure 7. (a). Schematic diagrams of CO2-induced reconfiguration on SYWC electrode surface. (b). Arrhenius plots of the SSNF cathode and surface reconstruction of SSNF. Reproduced with permission from Ref. [101].
Figure 7. (a). Schematic diagrams of CO2-induced reconfiguration on SYWC electrode surface. (b). Arrhenius plots of the SSNF cathode and surface reconstruction of SSNF. Reproduced with permission from Ref. [101].
Processes 13 03931 g007
Table 1. Representative doped SOFC cathodes under CO2 suggests that a higher valence improve apparent CO2 tolerance via stability.
Table 1. Representative doped SOFC cathodes under CO2 suggests that a higher valence improve apparent CO2 tolerance via stability.
MaterialTemperature (°C)CO2 (vol%)Time
(h)
Impedance: Start to End
BSCF7001
3
50Relative value: 1.0 to 2.8
Relative value: 1.0 to 12
BSCF-10%Y
(BSCF10Y)
7001
3
50Relative value: 1.0 to 1.7
Relative value: 1.0 to 8
BCFZY7001
10
120Rp: 0.368 to 0.574 Ω cm2
Rp: 0.503 to 0.682 Ω cm2
BCaCFZY7001
10
120Rp: 0.357 to 0.423 Ω cm2
Rp: 0.431 to 0.495 Ω cm2
PBCC75011001Rp: ≈0.024 Ω cm2
SSTF75600100.5Relative value: 1 to ~8.2
LCFe7001072Rp: ~0.24 to 0.29 Ωcm2
Table 2. Stability of representative doped SOFC cathodes under CO2 before and after adding SDC/GDC.
Table 2. Stability of representative doped SOFC cathodes under CO2 before and after adding SDC/GDC.
MaterialTemperature (°C)CO2 (vol%)Time (h)Impedance: Start to End (Ω cm2)
BLF (Ba0.95La0.05FeO3−δ)7003
(+5% H2O)
24Rp: 0.161 to 0.773
BLF-GDC (mechanical mixture)7003
(+5% H2O)
24Rp: 0.148 to 0.274
BLF-GDC (nano-GDC surface coating)7003
(+5% H2O)
24Rp: 0.126 to 0.160
BLF700102Rp: 0.190 to 0.264
BLF-30% SDC (biphasic)700102Rp: 0.121 to 0.150
SCNT (SrCo0.8Nb0.1Ta0.1O3−δ)5501100Rp: ~0.13 to ~1.3
SCNT-GDC composite5501100Rp: ~0.17 to ~0.8
Table 3. Stability of representative doped SOFC cathodes under CO2 before and after adding infiltration and coating.
Table 3. Stability of representative doped SOFC cathodes under CO2 before and after adding infiltration and coating.
MaterialTemperature (°C)CO2 (vol%)Time
(h)
Impedance: Start to End (Ω cm2)
LSCF (blank)7008100Rp: 0.26 to 1.92
LSCF-BaCoO3−δ coating
(BCO-LSCF)
7008100Rp: 0.11 to 0.32
PBSCF (PrBa0.5Sr0.5Co1.5Fe0.5O5+δ)700100.167
(10 min)
Rp: 2.50 to 3.63
PBSCF-LN
(La2NiO4+δ-coated PBSCF)
7001020Rp: 1.19 to ~1.56
Table 4. Comparative strategies for CO2-tolerance enhancement.
Table 4. Comparative strategies for CO2-tolerance enhancement.
StrategyMechanismOperation ComplexityAdvantagesLimitations
High-acidity/high-valence cation dopingIncreases ABE and lowers carbonate-formation ΔG; intrinsically stabilizes lattice/surfaceModerateIntrinsic and durable stabilization; scalable with standard ceramic processing; improved long-term CO2 tolerancePossible conductivity/ORR trade-off; requires optimization of sintering, phase compatibility and TEC; some dopants add cost
Composite cathodes (e.g., SDC/GDC)Dilutes Ba/Sr exposure and supplies stable ORR sites; microstructural/interface buffering against CO2SimpleDrop-in friendly (mixing/light infiltration); fast, tunable gains via phase ratio/porosity control; maintains ORR while improving CO2 robustnessMust ensure phase/TEC compatibility; potential interdiffusion/interface reactions over long operation if co-fired
Surface infiltration/coatingLSM/LNO/BCO nano-layers tune adsorption/exchange and transport; shield susceptible sitesModerateRetrofittable to existing electrodes; fine surface control; targeted suppression of CO2 adsorption/carbonatesPerformance hinges on thickness/uniformity/durability; excessive layers add resistance
CO2-induced in situ self-reconstructionControlled CO2 forms reversible carbonate/oxide nano-films or core-shell; adaptive passivation with potential mid-T ORR boostComplexAdaptive protection without foreign phases; reversible performance rebound; possibility of performance enhancementNarrow process window (time/temperature/ P C O 2 ); over-accumulation to poisoning; higher demands on reproducibility and in-process control; strict material selectivity/compatibility required
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, F.; Luo, Q.; Sun, M.; Song, Z.; Li, J.; Chen, B.; Zhang, Y. CO2 Poisoning of Solid Oxide Fuel Cell Cathodes: Mechanisms, Solutions, and Progress. Processes 2025, 13, 3931. https://doi.org/10.3390/pr13123931

AMA Style

Liu F, Luo Q, Sun M, Song Z, Li J, Chen B, Zhang Y. CO2 Poisoning of Solid Oxide Fuel Cell Cathodes: Mechanisms, Solutions, and Progress. Processes. 2025; 13(12):3931. https://doi.org/10.3390/pr13123931

Chicago/Turabian Style

Liu, Fang, Quan Luo, Meishen Sun, Zhaoqi Song, Junbiao Li, Bin Chen, and Yuan Zhang. 2025. "CO2 Poisoning of Solid Oxide Fuel Cell Cathodes: Mechanisms, Solutions, and Progress" Processes 13, no. 12: 3931. https://doi.org/10.3390/pr13123931

APA Style

Liu, F., Luo, Q., Sun, M., Song, Z., Li, J., Chen, B., & Zhang, Y. (2025). CO2 Poisoning of Solid Oxide Fuel Cell Cathodes: Mechanisms, Solutions, and Progress. Processes, 13(12), 3931. https://doi.org/10.3390/pr13123931

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop