Open Access Article
Mohammadali Emadi
and
Eric Croiset
*
Department of Chemical Engineering, University of Waterloo, Waterloo, Ontario N2L 3G1, Canada. E-mail: smemadif@uwaterloo.ca; ecroiset@uwaterloo.ca
First published on 6th January 2026
Solid oxide electrolysis cells (SOECs) are a promising technology for CO2 electrolysis and subsequent conversion to useful chemicals. This paper presents the development of a new cathode material to enhance the performance of SOECs for CO2 electrolysis. The focus was on the cathode material of the SOEC since it is the limiting factor for CO2 electrolysis. Sr2Fe1.5Mo0.5O6−δ (SFM) has attracted much attention due to its decent performance for CO2 electrolysis. To enhance the SFM performance, it was modified by doping bismuth and nickel to make a new composition of Bi0.1Sr1.9Fe1.4Ni0.1Mo0.5O6−δ (BiSFMNi). The Ni-doping made it possible for Fe–Ni nanoparticles to exsolve in situ when the material was reduced by 5% H2/Ar. Structural characterization techniques like XRD showed that, during exsolution, the material changed from a pure double perovskite structure to a mixed-phase material with both Ruddlesden–Popper (RP) and residual double perovskite phases and metallic nanoparticles. Electron microscopy (SEM/TEM/EDS) showed that Ni migrated to the surface of the perovskite bulk where it forms Fe–Ni nanoparticles. This material was then used as the cathode of SOECs, and the results showed that these exsolved Fe–Ni nanoparticles improved the electrocatalytic activity of the CO2 reduction reaction (CO2RR). The fabricated cell achieved a current density of 1.3 A cm−2 at 800 °C under an applied voltage of 1.6 V, while it was 1.0 A cm−2 for the non-exsolved nanoparticle sample.
Due to the slower rate of CO2 electrochemical reduction in the SOEC cathode compared to the anode's oxygen evolution process,2 the system has a high resistance. The performance of many cathode materials is also poor due to their low CO2 catalytic activity and adsorption capacity. Therefore, the main objective of this research is to enhance the performance of CO2 reduction in SOECs. It is necessary to develop cathode materials with an optimal combination of high conductivity, structural stability, high CO2 adsorption capacity, and high electrocatalytic activity for CO2 splitting.
Nickel, platinum, and other precious metals have been studied for their use as fuel electrode materials, but they face limitations such as thermal instability, corrosion at high temperatures, and/or high cost or scarcity. These issues lead to performance degradation, making pure metals less ideal for long-term SOEC operation.3 To overcome these challenges, solid oxide cells often use alloys or composite materials that combine the benefits of different metals while minimizing their drawbacks.4 State-of-the-art electrodes are usually made of cermet, typically Ni with YSZ (Yttria Stabilized Zirconia).5 However, cermet cathodes such as Ni–YSZ are prone to coke deposition during CO2 electrolysis.6,7 As such, there have been efforts to develop alternative cathode materials like perovskites.
Perovskite materials possess mixed ionic-electronic conductivity and good chemical stability. Among perovskites that have been used in SOECs, Sr2Fe1.5Mo0.5O6−δ (SFM) has attracted much attention.8 SFM has notable conductivity and redox stability, and it is frequently recommended as an electrode for CO2 electrolysis in SOECs.8 While SFM is thermally and chemically more robust, especially for CO2 electrolysis, its catalytic activity, CO2 adsorption capability, and long-term phase stability still need to be enhanced, especially when compared to the conductivity and catalytic strength of Ni-based cathodes.9 Researchers have explored a variety of techniques recently to increase its electrocatalytic activity for CO2 electrolysis, including infiltration technique, doping, and in situ exsolution.
Doping SFM with foreign elements to change the defective system and the crystal structure is another method to improve its electrical conductivity and electrocatalytic activity. B-site doping has been well explored in the literature due to successful improvement in cell performance.10–12 A-site doping, however, has not been thoroughly researched, and a limited number of research papers were found on A-site doping of SFM perovskite for CO2 electrolysis. A-site doping can influence the crystal structure, oxygen vacancy concentration, and overall stability of the perovskite, thereby impacting its ionic and electronic conductivity. Recent studies have tried to enhance the performance by doping rare-earth and alkaline-earth metals on the A-site. For example, Sun et al.13 investigated the performance of doping La on the Sr2Fe1.5Mo0.5O6−δ perovskite. The results showed that Sr1.9La0.1Fe1.5Mo0.5O6−δ with LSGM as the electrolyte could achieve a current density of 2.76 A cm−2 at 850 °C and 1.5 V, which is higher than that of non-doped SFM. This is because La-doping enhances the oxygen surface exchange rate as well as the bulk diffusion coefficient. Yang et al.14 studied the effect of doping bismuth on the A-site of SFM perovskite as the cathode for CO2 electrolysis. By partially replacing Sr with Bi, the modified SFM showed better CO2 adsorption and improved oxygen ion conductivity. The cell delivered a high current density of 1620 mA cm−2 at 1.6 V and 850 °C.
When a solid solution is formed by mixing two or more components, the atoms of each element are uniformly distributed within the crystal lattice. However, under certain conditions, such as cooling, compositional changes, or changes in temperature and pressure, one of the components may become insoluble, leading to the formation of a separate phase within the structure.15 The newly formed phase will often have a different composition, crystal structure, and physical properties compared to the original solid solution. This can lead to a range of interesting properties and applications, such as improved mechanical strength, enhanced catalytic activity, and increased conductivity.15 Exsolution of transition metal particles from the base double perovskite cathode can lead to a promising electrode with enhanced CO2 catalytic activation.16,17 Li et al.18 studied the effect of adding Ni–Fe nanoparticles into SFM-SDC uniformly. The chemical adsorption of CO2 and the kinetics of its surface reactions were both improved by the exsolved NiFe nanoparticles. Under a 1.5 V applied voltage at 800 °C, SOECs using this unique cathode have shown a current density of 2.16 A cm−2 and consistent CO2 electrolysis performance for 500 hours. To effectively accelerate the CO2RR in SOECs, in situ exsolved Fe–Ni nanoparticles on a Sr2Fe1.35Mo0.45Ni0.2O6 (SFNM) double perovskite substrate (Fe–Ni@SFNM) have been developed by Lv et al.19 The SOEC with the Fe–Ni@SFNM-GDC (Gd0.2Ce0.8O1.9) cathode exhibits great stability and no coke deposition for 40 h at 1.2 V, as well as a current density of 0.9 A cm−2 at 1.6 V and 800 °C. During the synthesis procedure, nickel was added to the B-site of Sr2Fe1.5Mo0.5O6 to create a stoichiometric Sr2Fe1.35Mo0.45Ni0.2O6 (SFNM) double perovskite oxide. Using a high-temperature reduction process, Fe–Ni alloy nanoparticles were exsolved in place and equally anchored on the perovskite surface, while the double perovskite kept its original structure. Lv et al.20 showed that metallic Co can be exsolved from the parent perovskite in a reducing environment, resulting in the creation of cation defects and oxygen vacancies. Metallic Fe may subsequently be exsolved to produce CoFe alloy nanoparticles in Sr2Fe1.35Mo0.45Co0.2O6−δ (SFMC), along with the structure changing from double to layered perovskite at 800 °C. At 600 °C, in an oxidizing environment, the spherical CoFe alloy nanoparticles are first converted to flat CoFeOx, then at 800 °C the structure changes back to double perovskite while being entirely dissolved back into the bulk. Due to the synergistic impact of CoFe alloy nanoparticles and plentiful oxygen vacancies at the metal-oxide interfaces, the SFMC cathode decorated with CoFe alloy nanoparticles exhibits 1.2 A cm−2 at 1.6 V and 800 °C, which is 50% higher than the parent cathode of SFMC. Zhang et al.21 investigated the stability issued of exsolved nanoparticles on the SFM perovskite. Although the nanoparticles make the material perform better, they are unstable at high voltages (i.e., ≥1.6 V). Zhang et al. found out that the B-site vacancies are the primary cause of the degradation. Therefore, they supplemented the perovskite with external Fe. The results showed that the supplemented SFNM had better catalytic activity and less degradation. Xi et al.22 explored the exsolution of bimetallic nanoparticles of Fe–Cu on the SFM perovskite. This was done by doping Cu on the SFM structure and then performing reduction treatment. The results showed a current density of 2.5 A cm−2 at 1.5 V and 800 °C, which was higher than 1.6 A cm−2 of the unreduced electrode. Zhang et al.23 investigated the performance of double perovskite and layered perovskite of SFM with nanoparticles exsolved. By introducing a B-site iron supplement strategy, the authors successfully promote nanoparticle exsolution while preserving the perovskite scaffold, breaking the traditional trade-off. Double perovskite-based nanoparticles (DP-NPs) showed the highest activity, while layered perovskite-based nanoparticles (LP-NPs) offered superior stability. Hu et al.24 proposed a novel strategy of dual-exsolution of nanoparticles on a composite electrode made of SFM and GDC. The cell delivered 1.72 A cm−2 at 1.5 V for the CO2RR, which is almost double that of the SFM/GDC electrode with no nanoparticles. Liu et al.25 studied the effect of doping zinc on Fe-sites of SFM to improve the Fe nanoparticle exsolution and ultimately increase the CO2RR performance. The cell achieved a high current density of 2.74 A cm−2 at 850 °C and 1.6 V, while for SFM with no nanoparticles it was 1.42 A cm−2. Tan et al.26 presented a high-performance solid oxide electrolysis cell (SOEC) cathode composed of Pr0.4Sr1.6(NiFe)1.5Mo0.5O6−δ (PSNFM) double perovskite, decorated with exsolved NiFe/FeOx (NFA@FeO) core–shell nanoparticles. The in situ exsolved particles and generated oxygen vacancies significantly enhanced CO2 adsorption and reduction activity. As a result, the cell achieved a high current density of 1.58 A cm−2 at 1.4 V and 800 °C. López-García et al.27 explored a double perovskite, SrxFeCo0.2Ni0.2Mn0.1Mo0.5O6−δ (x = 2.0, 1.9, 1.8), as a cathode material for SOECs. It demonstrates the formation of tunable Fe–Co–Ni ternary alloy nanoparticles via in situ exsolution, influenced by reduction temperature and A-site stoichiometry. Exsolution at 600 °C resulted in smaller, more abundant nanoparticles and better CO2 electrolysis performance than higher-temperature treatments, which led to RP phase formation. The x = 2.0 composition showed superior nanoparticle formation and conductivity. These findings highlight the potential of tailored alloy nanoparticle exsolution for enhancing the SOEC efficiency and broader energy conversion applications.
Exsolution has emerged as a highly promising strategy for enhancing electrode performance in solid oxide electrolysis cells (SOECs). Unlike conventional infiltration methods, exsolved nanoparticles are socketed into the perovskite matrix, offering excellent thermal stability and strong metal–support interaction. This approach not only prevents particle agglomeration under harsh conditions but also introduces abundant active sites and oxygen vacancies, significantly boosting the catalytic activity and long-term durability. Although prior studies have shown that exsolved nanoparticles from perovskite cathodes can improve CO2 electrolysis performance, most efforts have focused on single-site modifications or base SFM structure. In contrast, our work introduces a dual-site doping strategy, incorporating Bi on the A-site and Ni on the B-site of the SFM double perovskite. This novel approach aims at enhancing catalytic activity through exsolved Ni nanoparticles and also at leveraging Bi doping to improve structural stability and CO2 adsorption.
:
citric acid
:
total metal ions was kept at 1
:
1.5
:
1. NH3H2O was added to adjust the pH value to around 8. The solution was stirred and heated on a hot plate at 80 °C until the gel formed. The gel was then heated on a ceramic dish inside an oven at 400 °C for 4 hours for combustion and removing the organic components and the nitrates. After combustion, the porous material was ground to obtain powder. The powder was sintered at 1100 °C (temperature ramp rate of 3 °C min−1) for 5 hours in air to obtain the final pure phase perovskite powder. La0.4Ce0.6O2−δ (LDC) powder was prepared with the same procedure for BiSMFNi except that LDC was sintered at 1000 °C. LSGM powder and LSCF/GDC paste were purchased from fuelcellmaterials and used without further modification.
:
2 weight ratio of powder to ink. The resulting paste was then spin-coated onto the LSGM substrate and sintered at 1200 °C (temperature ramp of 3 °C min−1) for 5 hours to form a buffer layer prior to deposition of the BiSFMNi cathode. For the air electrode, a commercial LSCF/GDC paste (La0.6Sr0.4Co0.2Fe0.8O3−δ mixed with GDC) was purchased from fuelcellmaterials. The paste was used as received without further modification and directly applied onto the electrolyte substrate by screen printing. To make the cathode paste for screen printing, the synthesized BiSFMNi powder was mixed with a commercially available ink vehicle (fuelcellmaterials) in a 1
:
1 weight ratio and ball milled to make a paste that was smooth and easy to print. The paste was used to put the cathode layer on top of the LDC layer using a screen printer with an area of 1 cm2. After screen-printing the electrodes, the cell was sintered at 1100 °C (temperature ramp of 3 °C min−1). The overall cell fabrication process is shown in Fig. 1.
:
1 weight ratio, pressed into pellets, and sintered at 1100 °C for 5 hours. XRD analysis was then performed to evaluate phase stability. As shown in Fig. 3(b), the XRD patterns of BiSFMNi–LDC and LDC–LSGM correspond to the simple superposition of the individual phases, with no evidence of additional diffraction peaks, indicating strong chemical compatibility. A small amount of new phase, at 30° and 35°, in the BiSFMNi–LSGM combination suggests some interfacial interaction between these two materials. These results confirm the need of a thin LDC buffer layer to reduce unwanted interfacial interactions and enhance the chemical stability between the LSGM electrolyte and the BiSFMNi cathode.
XRD analysis was conducted on BiSFMNi samples under reducing conditions to investigate the exsolution behavior; the results are plotted in Fig. 3(c). The oxidized version of BiSFMNi showed a pure double perovskite phase. After 1 hour of reduction in 5% H2/Ar at 800 °C, however, the XRD pattern showed a mix of leftover double perovskite and newly formed secondary phases of a Ruddlesden–Popper (RP) structure and metallic Fe–Ni. This means that the perovskite scaffold breaks down in part and nanoparticles start to exsolve. The Fe–Ni peaks got stronger as the reduction time increased, while the double perovskite peaks got weaker. This shows that the structure is transforming and that the RP phase is becoming more dominant. After 5 hours of reduction, the XRD pattern predominantly showed the Ruddlesden–Popper phase along with Fe–Ni alloy peaks. This suggests that the original double perovskite structure had largely transformed, and a significant number of nanoparticles had been exsolved. Fig. 3(d) shows the Rietveld refinement that was performed on the XRD pattern of the BiSFMNi sample that had been reduced for one hour. The sample had 6.56% Fe–Ni alloy, 27.61% DP phase, and 65.83% RP phase.
The chemical stability of BiSFMNi was evaluated under CO/CO2 atmospheres at 800 °C. Powder and pellet samples were exposed for 24 h in either pure CO2 or 50% CO/CO2, followed by XRD and Raman analyses. As shown in Fig. S2, BiSFMNi retained its double perovskite structure in both atmospheres, with only weak SrCO3 peaks (∼25°) appearing under 50% CO/CO2 conditions, suggesting minor surface reactions. Raman spectra showed no D (∼1340 cm−1) or G (∼1580 cm−1) bands, confirming the absence of carbon deposition. These results demonstrate that BiSFMNi possesses excellent phase stability and carbon tolerance in CO/CO2 environments, supporting its suitability as a durable SOEC cathode material.
![]() | ||
| Fig. 4 XPS spectra of (a) Fe 2p, (b) Ni 2p, (c) Mo 3d, and (d) O 1s regions before and after 1 h reduction in 5% H2–Ar. | ||
![]() | ||
| Fig. 5 SEM images of BiSFMNi in three states: as-synthesized, reduced in 5% H2/Ar at 800 °C for 2, 10, 30 and 60 minutes, and reoxidized in air. | ||
The exsolution behavior and structural evolution of BiSFMNi were investigated using transmission electron microscopy techniques. Fig. 6 shows TEM images of the sample before and after reduction at 800 °C for 2 hours in a 5% H2/Ar atmosphere. More details of TEM analysis are presented in Fig. S4 and S5. In the as-synthesized state, the perovskite exhibits a cubic, dense surface morphology without any apparent nanoparticle decoration. After reduction, clear morphological changes are observed. Numerous nanoparticles emerged and were uniformly distributed on the surface of the perovskite matrix. In the Supplementary Materials, more details of TEM analysis and elemental composition are provided. Additionally, EDS elemental mapping was performed to provide a clearer visualization of the spatial distribution of individual elements across the sample. This technique helped distinguish the regions containing exsolved nanoparticles from the perovskite and Ruddlesden–Popper phases based on elemental composition. The EDS maps, shown in Fig. 6(c), show the spatial distribution of key elements: Fe and Ni signals are strongly localized in the exsolved nanoparticle regions, confirming the formation of Fe–Ni alloy nanoparticles, consistent with Fe–Ni, as identified by XRD analysis. In contrast, Sr, Mo, Bi, and O signals are concentrated within the bulk matrix, indicating their retention in the perovskite lattice. The presence of Bi remains uniformly distributed in the matrix, suggesting that Bi remains stable during reduction and does not participate in nanoparticle formation.
Overall, the combination of SEM, TEM, and EDS analyses confirms that the reduction treatment induces selective exsolution of B-site elements (Fe and Ni), while the A-site dopant (Bi) and other structural components remain integrated in the perovskite lattice.
The SEM cross-sectional image of the SOEC in Fig. 7 shows a well-defined multilayer structure, confirming successful layer deposition and good interfacial contact between components. The LSGM electrolyte layer is around 275 µm thick and gives the structural strength and ionic conductivity. On the fuel side (bottom), there is a 40 µm-thick porous cathode (BiSFMNi) that is separated from the electrolyte by a 5 µm LDC buffer layer. This thin layer was deposited via spin coating, which proved to be highly effective for achieving a uniform and continuous coating. The LDC layer makes the cathode and the LSGM electrolyte more chemically compatible and stops unwanted reactions at the interface. There is a 55 µm-thick porous anode (LSCF/GDC) on the oxygen side (top). The microstructure of both electrodes is porous, which is necessary for gas diffusion and offering a triple-phase boundary during CO2 electrolysis. The clear layer boundaries and the strong interfaces show that the screen-printing and sintering procedures worked well. Line EDS analysis across the SOEC cross-section (Fig. S6) confirmed distinct compositional layers with sharp interfaces. The LDC buffer layer was identified from elevated La and Ce signals between the LSGM electrolyte and the BiSFMNi cathode. Strong Fe, Mo, Bi, and Ni signals verified the BiSFMNi layer, and no significant elemental interdiffusion was observed, confirming the effectiveness of the LDC layer in preventing undesired reactions.
From thermal expansion coefficient (TEC) measurements (shown in Fig. S7), the average CTE values for SFM and BiSFMNi in the temperature range of 25–800 °C are 15.3 × 10−6 K−1 and 13.9 × 10−6 K−1, respectively. The noticeable decrease in TEC after Bi and Ni incorporation indicates that the dopants effectively decrease lattice expansion. This reduction in thermal expansion improves the compatibility of the cathode with typical solid oxide electrolytes. For comparison, LSGM electrolytes exhibit a TEC value of 11.4 × 10−6 K−1, while La-doped ceria (LDC), commonly used as an interlayer material, shows a TEC of 13.4 × 10−6 K−1. Therefore, BiSFMNi displays a closer thermal match to both LSGM and LDC than pristine SFM, effectively reducing the thermal strain at the cathode–electrolyte interface during high-temperature operation.
Fig. 8(b) shows the relative improvement in the current density of Fe–Ni@BiSFMNi versus BiSFMNi as a function of applied voltage. This was done to measure how much better the performance was with Fe–Ni nanoparticle exsolution. The data clearly show that increasing the voltage led to an increase in current density. The Fe–Ni@BiSFMNi cathode always outperformed the as-prepared BiSFMNi one over the entire voltage range. At a lower voltage around 1–1.2 V, the improvement in the current density is due to reduction in the activation overpotential. However, this improvement is small as the activation overpotential has a small contribution to the total overpotential. But when the voltage goes above 1.4 V, the improvement gets larger, reaching over 50% at 1.8 V. These results support the idea that exsolution is beneficial to increase the performance of SOEC cathodes, especially when they are under high load or demand.
The Fe–Ni@BiSFMNi-based cell was considered for further analysis. The effect of temperature between 700 and 800 °C on the I–V curve was investigated and plotted in Fig. 8(c). As expected, increasing the operating temperature led to increased current densities at all voltages. This improvement can be attributed to faster electrochemical reaction kinetics, especially for the CO2 reduction process, which is very thermally dependent.28 At 800 °C, the cell exhibits the highest performance, with a current density exceeding 1.3 A cm−2 at 1.6 V. In contrast, at 750 °C and 700 °C, both activation and ohmic losses are more pronounced, leading to a significantly lower current response of 0.67 A cm−2 and 0.23 A cm−2, respectively.
Constant-voltage tests were performed at 1.2 V, 1.3 V, and 1.4 V for 20 minutes at 800 °C to see how stable the Fe–Ni@BiSFMNi cathode was in the short term. Fig. 8(d) shows that all three curves start with a drop in current density in the first few seconds. After this change, the current density remains stable, which suggests that the electrode retains its structural and electrochemical integrity throughout the test.
Fig. 9(a) shows the EIS spectra of the BiSFMNi and Fe–Ni@BiSFMNi cathodes at 800 °C at different voltages (OCV, 1.2 V, and 1.4 V). As the voltage rises, the arc diameter clearly gets smaller (lower Rp) and also shifts a bit to a lower resistance (lower Rs). This pattern shows that charge transfer and catalytic activity are better at higher operating voltages. At 1.4 V, Rs equal 0.2 Ω cm2 and 0.21 Ω cm2 for BiSFMNi and Fe–Ni@BiSFMNi, respectively. Fe–Ni@BiSFMNi has smaller arcs (Rp) at all working voltages, which confirms that the exsolved Fe–Ni nanoparticles lower the polarization resistance, from 0.38 Ω cm2 for BiSFMNi to 0.30 Ω cm2 at 1.4 V, by increasing the number of reaction sites on the surface and improving the catalytic efficiency. Fig. 9(b) shows how temperature affects the polarization resistance of the Fe–Ni@BiSFMNi cathode at 1.2 V. The size of the arc gets larger resulting in an increase of the Rp from 0.45 Ω cm2 to 3.6 Ω cm2 (at 1.2 V) when the temperature decreased from 800 to 700 °C. This shows that higher temperatures favor CO2 reduction. These EIS results show that nanoparticle exsolution and higher operating temperatures both help reduce polarization losses and make the SOEC work better overall.
Traditional electrochemical impedance spectroscopy (EIS) is useful for measuring the total resistance of a cell, but it doesn't always have enough precision to tell among overlapping electrochemical processes. To get around this problem, the Distribution of Relaxation Times (DRT) analysis is used. DRT breaks down the impedance spectrum into separate time-constant parts, making it easier to find and assign resistive and capacitive processes in the cell. This method is very good at breaking down contributions from electrode polarization, gas diffusion, and charge transfer at the interface.
The DRT deconvolution, Fig. 9(c), shows four separate peaks, which are linked to electrochemical processes that happen at various characteristic frequencies:
• P1 and P2: attributed to oxygen ion transfer in the electrolyte and at interfaces and oxygen evolution at the anode.19,22,23,29
• P3: associated with surface processes such as CO2 adsorption, activation, and electrochemical reduction.19,22,23
• P4: gas diffusion process.19,22,23
It is observed that P1, P2, and P4 do not change significantly between samples, suggesting that they predominantly reflect bulk ionic transport and gas diffusion phenomena, which are less sensitive to surface modifications. In contrast, P3 shows a more pronounced change, indicating that it is closely related to CO2 reduction processes at the cathode surface. The smaller P3 peak for Fe–Ni@BiSFMNi demonstrates that the exsolution of Fe–Ni nanoparticles accelerate CO2 surface kinetics by enhancing adsorption and activation. The exsolution achieved through reduction treatment increases the number of metal-oxide interfaces and oxygen vacancies, both of which facilitate gas-phase reactions and improve electrochemical surface exchange rates. These results are consistent with the I–V and EIS studies, which showed improved current density and reduced polarization resistance.
The DRT profiles also show the effect of temperature on the performance, as shown in Fig. 9(d). As the temperature drops, all the peaks move to lower frequencies and are stronger, which means that the kinetics are slower, and the resistive losses are larger. The general pattern is that the polarization resistance increases significantly as the temperature goes down, especially in the low-frequency range where CO2 reduction happens.
000 hours, the cell maintains relatively high performance over the 140 hour test. The gradual current decay (∼0.6 mA h−1) suggests that, if continued linearly, the cathode would not meet practical lifetime requirements. However, the current may stabilize over extended operation due to early-stage electrode restructuring, and further long-term studies are required to confirm stable performance over practical timescales.
![]() | ||
| Fig. 10 (a) Long-term CO2 electrolysis performance at 1.3 V over 140 hours and (b) Raman spectra of the surface of the Fe–Ni@BiSFMNi cathode after the durability test. | ||
Raman spectra collected from the BiSFMNi cathode surface before and after 140 h of CO2 electrolysis are shown in Fig. 10(b). No characteristic D (1340 cm−1) or G (1580 cm−1) bands associated with carbon species were detected, confirming that no coke formation occurred during long-term operation. The stability of the Raman profile further supports the high tolerance of the BiSFMNi electrode toward carbon deposition and its good catalytic durability during CO2 electrolysis.
Supplementary information (SI): additional characterization data supporting the main text, including XRD comparisons of as-synthesized, reduced, and reoxidized BiSFMNi; CO/CO2 stability tests; thermal stability results; and thermal expansion coefficient tests. See DOI: https://doi.org/10.1039/d5ta07206a.
| This journal is © The Royal Society of Chemistry 2026 |