Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Perovskite-triggered dual exsolution of oxygen-deficient CeO2 matrix and NiFe nanoalloys for enhanced CO2 electrolysis

Biao Ouyang a, Ming Yang a, Lin-Bo Liu a, Shuo Liu a, Yan Li a, Xian-Zhu Fu b, Yifei Sun cd, Subiao Liu *a and Jing-Li Luo be
aSchool of Minerals Processing and Bioengineering, Central South University, Changsha, Hunan 410083, China. E-mail: subiao@csu.edu.cn; Tel: +86 15675160502
bCollege of Materials Science and Engineering, Shenzhen University, Shenzhen 518055, China
cCollege of Energy, Xiamen University, Xiamen 361005, China
dShenzhen Research Institute of Xiamen University, Shenzhen 518000, China
eDepartment of Chemical and Materials Engineering, University of Alberta, Edmonton, Alberta T6G 1H9, Canada

Received 20th August 2025 , Accepted 16th October 2025

First published on 17th October 2025


Abstract

Perovskite cathodes for CO2 electrolysis offer excellent redox stability but suffer from limited activity. Although in situ exsolution of B-site cations is a powerful strategy to alleviate this issue, this process often triggers co-segregation of insulating AOx phases, which diminishes active site exposure and ionic-electronic conductivity. Here we addressed the critical issue of insulating phase segregation in conventional exsolved perovskite by developing a Ce-doping strategy in Sr1.95Ce0.05Fe1.3Ni0.2Mo0.5O6−δ (SCeFNM). This material, upon annealing in a reducing atmosphere, allowed in situ construction of nanoscale CeO2–NiFe/oxide heterostructures (CeO2–NiFe@SCeFNM) by co-exsolving oxygen-deficient CeO2 phase and NiFe alloy nanoparticles (NPs) on the surface. The unique architecture achieved a high current density of 1.57 A cm−2 at 1.5 V and 850 °C with a CO faradaic efficiency of over 96%, outperforming NiFe@SFNM and their counterparts. Combined results demonstrated that the superior activity mainly came from the synergy within the “three-in-one” heterostructure, where NiFe alloy NPs increased the electronic conductivity, and CeO2 phase extended the O2− migration channels capable of enhancing CO2 adsorption and activation, while the perovskite backbone ensured structural integrity. This study establishes a universal paradigm for constructing advanced catalysts for diverse applications via co-exsolution of metal NPs and defective oxide phases from appropriately A-site-doped perovskites.



Green foundation

1. High-temperature CO2 electrolysis within SOECs has emerged as a transformative pathway, which enables direct conversion of CO2 into CO or syngas using renewable energy. This study innovatively addresses the critical issue of an insulating SrO phase inherent in conventional exsolution processes by strategically constructing heterostructured CeO2–NiFe@SCeFNM as a SOEC cathode material, thereby advancing green chemistry and contributing to sustainable technologies.

2. We successfully achieved efficient conversion of CO2 into industrially vital CO using renewable electricity in an SOEC, which delivered a high current density of 1.57 A cm−2 at 850 °C and 1.5 V, together with a high CO faradaic efficiency of over 96%, therefore significantly enhancing carbon utilization efficiency.

3. Future optimization to avoid AO phase segregation of perovskites via lowering operating temperatures or substituting rare-earth elements at A-sites could further advance their sustainability and scalable implementation of electrocatalytic carbon-neutral technologies.


1. Introduction

The increasing consumption of fossil fuels has led to significant carbon dioxide (CO2) emissions, which profoundly impact global warming. An urgency to address climate change, therefore, has propelled CO2 reduction technologies to the forefront of sustainable energy research, offering a dual solution for CO2 utilization and renewable energy storage.1,2 Among them, high-temperature CO2 electrolysis within solid oxide electrolysis cells (SOECs) has emerged as a promising technology, enabling direct conversion of CO2 into CO or syngas using renewable energy.3,4 Despite the remarkable catalytic performance of Ni-yttrium stabilized zirconia (Ni-YSZ) cermet as an SOEC cathode, it suffers from intrinsic redox instability, agglomeration and severe coking issues, which critically limit its further utilization at larger industrial-scale application.5–7 Alternatively, perovskite oxides (e.g., ABO3, A2BO4) have been extensively investigated as promising catalysts for SOECs due to their mixed ionic-electronic conductivity and superior redox stability.8,9 Nevertheless, a major limitation lies in their relatively inferior catalytic activity, which leads to an insufficient electrolysis performance, highlighting the imperativeness of exploring smart design strategies for perovskite oxide electrodes.

Surface engineering via in situ exsolution has emerged as a powerful strategy to enhance catalytic activity by enabling controlled growth of nanoscale metallic nanoparticles (NPs) on perovskite oxide surfaces.4,10–12 This strategy not only achieves a uniform dispersion of catalytic NPs across porous electrode architectures, but also generates abundant oxygen vacancies at the metal–oxide interfaces, thereby creating additional electrochemically active sites that improve CO2 adsorption and activation capabilities. Furthermore, it ensures a high operational stability by mitigating active site loss caused by catalyst detachment and/or agglomeration, while maintaining structural integrity under harsh operating conditions.13 Among various metal NPs (e.g., Fe, Co, Ni, and Cu), in situ exsolution of Fe-/Ni-based metals/alloys has been widely employed since Fe shows exceptional CO2 adsorption capacity and coking resistance, while Ni demonstrates a superior catalytic activity with the lowest activation energy barrier.12,14 However, the exsolution of B-site reducible cations may induce the intergrowth of AOx phases, which alters the functional properties of the material.11,15 Previous studies have demonstrated that exsolution of B-site cations from perovskite oxides triggers SrO co-segregation on occasion—a process driving chronic electrochemical performance decay since the electronically and ionically insulating nature of surface-terminated SrO obstructs surface charge transfer and oxygen transport pathways.16,17 Moreover, in cation-ordered LnBaCo2O5+δ perovskites (Ln = lanthanide), BaO segregation is also kinetically accelerated at temperatures exceeding 400 °C, resulting in a diminished surface coverage of Co active sites on the perovskite oxide framework.18

However, it is worth noting that the exsolution of A-site oxides is not always detrimental, since precise design of A-site oxide exsolution offers unprecedented opportunities to tailor CO2 adsorption and oxygen exchange reactions, which serves as an alternative approach to enhancing SOEC performance. For instance, the exsolution of CaO and NiFe NPs from La0.6Ca0.4Fe0.8Ni0.2O3−δ doubled its specific surface area.19 Likewise, Zhu et al. synthesized self-generated Fe/MnOx NPs by regulating A-site deficiency, and the interfacial charge redistribution of Fe/MnOx optimized CO2 dissociation pathways and promoted chemical adsorption and dissociation of CO2 molecules on the cathode surface.20 Moreover, the exsolution of oxides can also increase oxygen vacancy content and boost O2− migration. Previous studies have reported that Bi-doped La0.4Sr0.6Co0.2Fe0.7Nb0.1O3−δ exsolved Bi2O3 and CoFe alloy NPs, which raised the oxygen vacancy concentration by 40%,21 while oxygen vacancy-enriched PrOx NPs enhanced O2− migration capability, which accelerated O2− exchange and migration processes on the electrode surface.22 Therefore, precise engineering of oxide-metal/oxide heterointerfaces holds the potential to enhance CO2 electrolysis activity and stability. Cerium oxide (CeO2), due to its exceptional oxygen storage capacity and redox cycling between Ce3+ and Ce4+, possesses a desirable level of O2− conductivity, and the redox couple of Ce4+/3+ on the surface accelerates O2− transfer between CO2 and CO. Meanwhile, NiFe alloy NPs, characterized by their high electronic conductivity and robust thermal stability, are promising candidates for CO2 dissociation and CO desorption. However, individual NiFe alloy NPs suffer from insufficient O2− migration, while pure CeO2 lacks the electronic conductivity required for efficient charge transfer in CO2 electrolysis.

Recognizing this complementary potential between CeO2 and NiFe alloy NPs, herein we designed and synthesized a novel Ce-doped heterostructured catalyst by annealing Sr1.95Ce0.05Fe1.3Ni0.2Mo0.5O6−δ (SCeFNM) in a reducing atmosphere to strategically construct metallic NPs with oxygen-vacancy-rich CeO2 on SCeFNM (CeO2–NiFe@SCeFNM), realizing dual exsolution of NiFe alloy and CeO2 NPs with enhanced electronic conductivity, O2− migration, and CO2 adsorption and activation. Experimental results demonstrated that a cell with a CeO2–NiFe@SCeFNM cathode achieved a comparable current density (j) of 1.57 A cm−2 at 850 °C and 1.5 V for pure CO2 atmosphere, together with a high CO faradaic efficiency (FECO) of over 96% and a stability of more than 120 hours, showcasing an exceptional catalytic activity for CO2 electrolysis. The strategic integration of dual phases through doping-induced co-precipitation offers a compelling approach to synergizing their complementary functionalities.

2. Experimental methods

2.1 Material synthesis

Sr2Fe1.3Ni0.2Mo0.5O6−δ (SFNM), SCeFNM, and La0.4Ce0.6O2 (LDC) were obtained via a sol–gel method.23 Stoichiometric amounts of Sr(NO3)2, Fe(NO3)3·9H2O, Ni(NO3)3·6H2O, (NH4)6Mo7O24·4H2O and Ce(NO3)3·6H2O were dissolved in EDTA-NH3·H2O combined solution under continuous heating and stirring; citric acid was then introduced into the solution. The molar ratio of EDTA[thin space (1/6-em)]:[thin space (1/6-em)]citric acid[thin space (1/6-em)]:[thin space (1/6-em)]total cation was controlled to be around 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5[thin space (1/6-em)]:[thin space (1/6-em)]1. Subsequently, NH3·H2O 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 formation of organic resins containing homogeneously distributed cations because of the slow evaporation of the solvent. The formed gels were heated at 300 °C for four hours to remove the organic components and the nitrates, followed by being annealed at 1100 °C for 10 hours to achieve the target powders. NiFe@SFNM and CeO2-NiFe@SCeFNM were prepared by annealing SFNM and SCeFNM powders at 850 °C for 2 hours in a 5% H2/Ar atmosphere with a flow rate of 50 mL min−1 in a tubular furnace. The heating and cooling rates were both controlled at 2 °C min−1. The La0.8Sr0.2Ga0.8Mg0.2O3−δ (LSGM) electrolyte was prepared by a dry-press method, and then calcined at 1450 °C for 5 hours. The (La0.6Sr0.4)0.95Co0.2Fe0.8O3−δ (LSCF) and Ga0.2Ce0.8O2−δ (GDC) powders were used as received.

2.2 Material characterization

The crystalline structure of all the synthesized powders was identified by X-ray diffraction (XRD) with Rigaku Rota flex Cu Kα radiation (40 kV, 44 mA), and the raw data were analyzed with JADE version 6.5. Microstructures were determined with a high-resolution Zeiss Sigma field emission scanning electron microscope (SEM). The low- and high-resolution transmission electron microscopy (TEM) images were acquired by using a Thermo Scientific Talos F200X STEM at an accelerating voltage of 200/300 kV. Energy-dispersive X-ray spectroscopy (EDX) combined with high-resolution TEM (HR-TEM) was applied to characterize the composition and element distribution. X-ray photoelectron spectroscopy (XPS, Kratos AXIS Ultra) was used to investigate the surface chemistry of SFNM, SCeFNM, and CeO2–NiFe@SCeFNM with adventitious carbon (C 1s) at a binding energy (BE) of 284.8 eV as the reference. Temperature-programmed desorption of CO2 (CO2-TPD) was investigated with a Micromeritics chemisorption analyzer (Auto Chem II 2920) to study the CO2 desorption. The CO2 adsorption ability was also characterized by Fourier transform infrared (FTIR) spectroscopy (Nicolet iS50). The oxygen vacancy content was measured using an iodometric titration method under Ar atmosphere.

2.3 Cell fabrication and measurement

The LDC buffer layer was formed by screen-printing an α-terpinol/ethyl cellulose-bound slurry onto LSGM followed by 1350 °C calcination for 2 hours. Subsequently, CeO2–NiFe@SCeFNM-GDC, NiFe@SFNM-GDC, SCeFNM-GDC, SFNM, and LSCF-GDC powders were separately mixed with a binder composed of α-terpinol and ethyl cellulose ether to form their corresponding slurries. GDC was incorporated into all cathodes primarily to enhance the oxygen-ion conductivity and extend the triple-phase boundaries. The electrode slurry was screen printed on the LSGM electrolyte, followed by calcining at 1100 °C for 2 hours, to obtain a cell configuration of CeO2–NiFe@SCeFNM-GDC∥LDC∥LSGM∥LSCF-GDC. The effective area of the electrode was 0.5 cm2 and Ag paste was used as the current collector. The as-fabricated single cell was mounted on one end of an alumina support for electrochemical measurements, with the LSCF-GDC anode exposed to ambient air. Before testing, the CeO2–NiFe@SCeFNM cathode compartment was purged with 5% H2/Ar. Then, pure CO2 was fed to the cathode at a flow rate of 50 mL min−1 for testing. A schematic diagram of the high-temperature CO2 electrolysis setup is shown in Fig. S1. The electrochemical performance of the single cell was evaluated using an EnergyLab XM electrochemical workstation from OCV to 1.6 V at a scan rate of 20 mV s−1. Electrochemical impedance spectra (EIS) were acquired with an applied AC amplitude of 10 mV in the frequency range of 0.1 Hz to 1.0 MHz. The CO2 and CO contents of outlet gas were determined by online gas chromatography (Agilent 8860) using a thermal conductivity detector.

3. Results and discussion

The pristine perovskite oxides, SFNM and SCeFNM, were synthesized using a facile sol–gel method, as reported elsewhere.24,25 The XRD patterns of SFNM and SCeFNM powders after calcination at 1100 °C were detected without impure phase (Fig. S2), but slight shifts of all diffraction peaks toward higher angles were observed for SCeFNM, indicating lattice volume shrinkage due to the partial substitution of a smaller ionic radius of Ce3+ (1.14 Å) to Sr2+ (1.34 Å). After annealing SFNM and SCeFNM powders at 850 °C in 5% H2/Ar for 2 hours, the double perovskite oxide phase of SCeFNM was transformed into a Ruddlesden–Popper phase (Fig. 1a), together with dual exsolutions of A-site CeO2 phase and B-site NiFe alloy NPs (CeO2–NiFe@SCeFNM). In contrast, only NiFe alloy NPs were in situ exsolved from SFNM (NiFe@SFNM), accompanied with a slight phase transformation. To gain insights into the microstructural features, SFNM, SCeFNM, NiFe@SFNM, and CeO2–NiFe@SCeFNM powders were subjected to SEM analysis. It was seen that SFNM and SCeFNM specimens showed smooth network structures (Fig. S3) whereas substantial NiFe alloy NPs appeared on the reduced SFNM and SCeFNM. Different from NiFe@SFNM, SCeFNM with Ce doping possessed a greater amount of exsolved NiFe alloy NPs, alongside a minor quantity of CeO2 NPs (Fig. 1b). For further confirmation, HR-TEM was conducted, and it was found that lattice spacings of 0.202 and 0.313 nm were observed in HR-TEM images (Fig. 1c–d and S4) which corresponded to the (111) planes of CeO2 and NiFe alloy, respectively. Accordingly, the high-angle annular dark-field scanning TEM (HADDF-STEM) image and EDX elemental mapping of CeO2–NiFe@SCeFNM (Fig. 1e and f) demonstrated that the perovskite oxide surface was uniformly dispersed with CeO2 and abundant NiFe alloy NPs with an average size of 17.5 ± 3.2 nm, while other elements (e.g., Sr, Mo, O) were homogeneously distributed within the perovskite oxide matrix. Moreover, the exsolved NPs were firmly embedded on the CeO2–NiFe@SCeFNM substrate, indicating the presence of strong oxide-metal/oxide interactions at the heterointerfaces.
image file: d5gc04390e-f1.tif
Fig. 1 (a) XRD patterns of NiFe@SFNM and CeO2–NiFe@SCeFNM, together with an enlargement of the region at 29–30°. (b) SEM image of CeO2–NiFe@SCeFNM. HR-TEM images and corresponding FFT patterns of (c) NiFe NPs and (d) CeO2 NPs. (e) HADDF-STEM image of CeO2–NiFe@SCeFNM with (f) EDX mappings of all elements.

Surface chemical variation of the as-synthesized perovskite oxides was characterized via XPS. The XPS full survey spectrum exhibited a Ce element signal in the CeO2–NiFe@SCeFNM sample, further confirming the presence of Ce4+ (Fig. S5). Clearly, Fig. 2a and b showed that an increase in A-site Ce-doping led to a decrease in the intensities of Fe3+ and Ni3+ species instead of an increase in those of Fe2+ and Ni2+ species. This implied that increasing A-site Ce-doping reduced the oxidation states of Fe and Ni, and caused an enhanced electronegativity, which induced the formation of more oxygen vacancies in turn to maintain the charge neutrality. Notably, two additional peaks at 707 and 852 eV were observed for CeO2–NiFe@SCeFNM, which were ascribed to the presence of Fe0 and Ni0. Moreover, the O 1s spectra were deconvoluted into three subpeaks, representing the lattice oxygen (OLatt), the adsorption oxygen (OAds), and the surface carbonate/hydroxyl oxygen (CO32−/OH) species, respectively.26 Typically, the peak of OAds represented oxygen vacancies, and it is worth noting that CeO2–NiFe@SCeFNM demonstrated a higher concentration of OAds as compared with SFNM and SCeFNM (Fig. 2c), confirming its oxygen vacancy concentration had a considerable enhancement, which aligned well with the iodine titration results in Fig. 2d, confirming its superiority for CO2 adsorption. To examine the impact of oxygen vacancies on CO2, its adsorption behaviors on SFNM, SCeFNM, and CeO2–NiFe@SCeFNM powders were explored in the temperature range of 50–850 °C using CO2-TPD. There were no desorption peaks observed at low temperatures (<300 °C) in Fig. 2e, while the peaks in the high-temperature range (>300 °C) were mainly associated with the desorption of chemisorbed CO2.27 Clearly, the peak area for CeO2–NiFe@SCeFNM, benefitting from the exsolved CeO2 phase with abundant oxygen vacancies, was significantly larger than those for SFNM and SCeFNM, indicative of a superior CO2 adsorption capacity.28 The elevated desorption temperature of 754 °C indicated an improved bonding ability between CO2 molecules and perovskite oxides,16 which coincided well with the working temperatures of SOECs. Furthermore, since carbonates are critical intermediate species in the CO2 electrolysis process,29 CO2 chemisorption was further characterized by FTIR spectroscopy with samples being exposed to CO2 at 800 °C for 10 hours. As shown in Fig. 2f, the CO32− absorption signals in the infrared region of ∼1440 cm−1 for CeO2–NiFe@SCeFNM were stronger as compared to SFNM and SCeFNM, confirming that CeO2–NiFe@SCeFNM could significantly enhance CO2 adsorption capacity attributed to the greater amount of oxygen vacancies.30 In particular, in situ FTIR spectra for CeO2–NiFe@SCeFNM all presented strong CO32− signals from 750 to 850 °C (Fig. S6), demonstrating its robust CO2 adsorption capability.


image file: d5gc04390e-f2.tif
Fig. 2 (a) Fe 2p spectra, (b) Ni 2p spectra, (c) O 1s spectra, (d) oxygen vacancy concentration determined by iodometric titration, (e) CO2-TPD plots and (f) FTIR spectra of SFNM, SCeFNM, NiFe@SFNM and CeO2–NiFe@SCeFNM.

To unravel the impacts of exsolved CeO2 phase and NiFe alloy NPs as well as their constructed oxide-metal/oxide heterointerface on CO2 electrolysis activity, SFNM, SCeFNM, NiFe@SFNM, and CeO2–NiFe@SCeFNM were fabricated as cathodes within SOECs supported by LSGM electrolyte, and were tested in pure CO2 atmosphere from 750 to 850 °C. Based on the current–voltage (IV) curves, a gradual increase in j was observed for all the samples with an order of SFNM < SCeFNM < NiFe@SFNM < CeO2–NiFe@SCeFNM (Fig. 3a and S7), delivering a j of 1.57, 1.45, 1.35 and 1.23 A cm−2 at 1.5 V and 850 °C, respectively (Table S1). This implied that as compared to conventional SFNM with only NiFe alloy NPs and pristine SFNM, CeO2–NiFe@SCeFNM provided more reactive centers precisely due to the dual exsolution of precipitated CeO2 and NiFe NPs, both of which synergistically facilitated the activity. Concurrently, the key reaction steps and kinetic behaviors during CO2 electrolysis were investigated using EIS (Fig. S8) and the distribution of relaxation times (DRT) (Fig. S9). The Nyquist plots in Fig. 3b illustrate that the polarization resistance (Rp) values of SFNM, SCeFNM, NiFe@SFNM, and CeO2–NiFe@SCeFNM cathodes under open-circuit voltage (OCV) at 850 °C were 1.025, 0.885, 0.857, and 0.792 Ω cm2, respectively, in accordance with the j variation trend. Given that the contributions from the LSCF anode and LSGM electrolyte to the total resistance were quite close for all cells, the polarization resistance discrepancy stemmed from the different cathodes. Based on previous studies involving DRT analysis it was known that the peaks at low, intermediate, and high frequency (LF, IF, and HF) are associated with CO2 adsorption and activation, migration of O2− in the electrodes, and charge/ion transfer at the electrode–electrolyte interface, respectively.13,31 Apparently, the dual exsolution treatment reduced the LF region area while leaving the IF and HF regions unaffected (Fig. 3c), suggesting that the diminished LF contribution for the CeO2–NiFe@SCeFNM cathode optimized the CO2 adsorption and activation kinetics. Furthermore, the Rp value of CeO2–NiFe@SCeFNM decreased markedly with increasing applied potential (Fig. S8a and Table S2), reflecting an accelerated CO2 electrolysis kinetics under high potentials. Moreover, increasing the applied potential could lower the LF region area, thus improving CO2 adsorption and activation, and the ultimate CO2 electrolysis activity (Fig. S9a). In contrast, raising the operating temperature lowered both the ohmic resistance (Ro) and Rp (Fig. S8e and Table S3), and the accordingly reduced IF and LF region areas in DRT plots revealed that the increase in temperature not only enhanced CO2 adsorption and activation, but also accelerated O2− migration at the electrode (Fig. S9e).32 This was further verified by the electrochemical activation energy, an important parameter to assess the CO2 electrolytic process. The Arrhenius plots in Fig. 3d show that the CeO2–NiFe@SCeFNM cathode delivered a smaller activation energy of 0.992 eV than its counterparts, indicative of a reduced cathodic energy barrier. As a result, benefiting from the synergy of NiFe alloy NPs and CeO2 phase, CeO2–NiFe@SCeFNM achieved the optimal CO2 electrolysis performance among all the control materials, and surpassed that of most of the state-of-the-art cathode materials reported to date, as shown in Fig. 3e, again confirming its superiority for CO2 electrolysis in SOECs.


image file: d5gc04390e-f3.tif
Fig. 3 (a) IV curves, (b) Nyquist plots and (c) DRT results of SFNM, SCeFNM, NiFe@SFNM and CeO2–NiFe@SCeFNM at 850 °C. (d) Arrhenius plots of various cathodes from 750 to 850 °C. (e) Activity comparison of all samples in this study with previously reported state-of-the-art materials for CO2 electrolysis at 1.5 V, where SR represents the reference in SI.

As crucial parameters to evaluate CO2 conversion efficiency, CO production rate (PRCO) and FECO are equally important for the development of SOECs. The potentiostatic CO2 electrolysis measurements at varying applied potentials revealed that CeO2–NiFe@SCeFNM achieved maximum j in a wide potential range from 1.0 to 1.5 V (Fig. 4a), which matched well with the IV results. Additionally, the corresponding FECO and PRCO under different applied potentials at 850 °C were also evaluated (Fig. 4b). The single cell with the CeO2–NiFe@SCeFNM cathode achieved a high FECO of 96.03%, surpassing the 90.94% on SFNM and closely matching those of 95.72% on NiFe@SFNM and 95.35% on SCeFNM in CO2-to-CO conversion, at 1.5 V and 850 °C. In addition, at 1.5 V, the PRCO of CeO2–NiFe@SCeFNM reached 10.32 mL min−1 cm−2, faster than those of 9.68, 8.68 and 8.42 mL min−1 cm−2 over NiFe@SFNM, SCeFNM, and SFNM, respectively, which further confirmed that except for j, co-exsolution of CeO2 phase and NiFe alloy NPs also improved CO2 conversion rate. The long-term stabilities of the cells with different cathodes were also investigated, and Fig. 4c shows j and FECO as a function of time under a constant potential of 1.2 V at 850 °C. The CeO2–NiFe@SCeFNM-assembled cell could maintain a good performance with a negligible j decay rate of 0.793 mA h−1 during continuous operation of 120 hours, and retained a steady FECO of 90%, superior to that of the cells with NiFe@SFNM, SCeFNM, and SFNM. For a better holistic view, a radar map presenting j (A cm−2), 1/Rp−1 cm−2), PRCO (mL min−1 cm−2), FECO and stability retention rate per 100 hours (Rsr = j100/jinitial)25 is plotted in Fig. 4d for comparison. It illustrated clearly that dual exsolution was more beneficial for CO2 electrolysis, which also highlighted the critical role of CeO2 phase in tuning CO2 electrolysis activity. In addition, no extra phase and impurity were observed in the XRD pattern post-test for the CeO2–NiFe@SCeFNM cathode (Fig. 4e), indicating a good structural stability under harsh electrolysis conditions. Meanwhile, the cathode-electrolyte interface remained closely stacked, and no delamination was detected after long-term CO2 electrolysis operation (Fig. 4f and S10a, b). Importantly, the perovskite oxide matrix still retained the uniformly anchored in situ exsolved NPs without severe agglomeration for the CeO2–NiFe@SCeFNM cathode (Fig. 4g). However, there were no noticeable in situ exsolved NPs on the NiFe@SFNM cathode (Fig. S10c), clarifying a good structural resilience under high-temperature harsh conditions in the mixture of CO2/CO, as also confirmed by the Raman spectrum of the CeO2–NiFe@SFNM cathode after stability testing in Fig. 4h. Notably, no obvious peaks associated with SrCO3 (1050 cm−1), D-bond (1340 cm−1) and G-bond (1580 cm−1) were observed, evidencing the good coking resistance and the strong suppression to Sr segregation of CeO2–NiFe@SCeFNM.27


image file: d5gc04390e-f4.tif
Fig. 4 (a) The potentiostatic measurements for CO2 electrolysis and (b) PRCO and corresponding FECO under different applied potentials of all samples. (c) Long-term stability tests and the corresponding FECO at 1.2 V and 850 °C for NiFe@SFNM and CeO2–NiFe@SCeFNM. (d) Radar map of SFNM, SCeFNM, NiFe@SFNM and CeO2–NiFe@SCeFNM. (e) XRD patterns and SEM images of (f) a cross-section of a single cell and (g) CeO2–NiFe@SCeFNM cathode as well as (h) Raman spectra of the CeO2–NiFe@SCeFNM cathodes before and after stability testing.

To clearly elaborate the mechanism credited for the superior electrochemical activity on CeO2–NiFe@SCeFNM, a schematic illustration of the possible reaction pathways over SFNM, NiFe@SFNM and CeO2–NiFe@SCeFNM is shown in Fig. 5. It was seen that when SFNM was assembled as a cathode, its activity was limited to the slow gas–solid interface exchange process involving CO2 adsorption, activation, and O2− migration on the cathode surfaces (Fig. 5a). Nevertheless, the introduction of NiFe alloy NPs characterized by high electronic conductivity could promote charge transfer, and CO2 adsorption and activation in CO2 electrolysis, but individual NiFe alloy NPs suffered from insufficient O2− migration (Fig. 5b). Thus, we proposed the “three-in-one” strategy in this study by integrating both NiFe alloy NPs and defective CeO2 phase in situ exsolved from SCeFNM (Fig. 5c), which synergistically increased the ionic-electronic conductivity and the O2− migration channels from CeO2 to SCeFNM to promote CO2 adsorption and activation. In addition, the intimate contact of SCeFNM, NiFe alloy NPs and CeO2 at the nanoscale also effectively extended the triple-phase boundary length dedicated to CO2 electrolysis, thus culminating in the exceptional CO2 electrolysis activity.


image file: d5gc04390e-f5.tif
Fig. 5 Schematic illustrations of possible reaction pathways for CO2 electrolysis over (a) SFNM, (b) NiFe@SFNM and (c) CeO2–NiFe@SCeFNM.

4. Conclusions

In summary, we synthesized Sr1.95Ce0.05Fe1.3Ni0.2Mo0.5O6−δ (SCeFNM) using a Ce-doping strategy, and strategically constructed CeO2–NiFe/oxide heterostructures by in situ co-exsolving oxygen-vacancy-rich CeO2 phase and NiFe alloy NPs from perovskite oxide (CeO2–NiFe@SCeFNM) to avoid the co-segregation of insulating SrO phase. It was found that CeO2–NiFe@SCeFNM cathode achieved a j of 1.57 A cm−2 at 850 °C and 1.5 V, and a CO faradaic efficiency of 96%, outperforming NiFe@SFNM without oxygen-deficient CeO2 phase, and pristine SFNM and SCeFNM. Crucially, the cathode demonstrated an exceptional operational stability over 120 hours without carbon deposition and particle agglomeration, overcoming the degradation limitations of traditional exsolution. Combined results and spectra collectively verified that the enhanced activity exhibited by CeO2–NiFe@SCeFNM was mainly derived from the three-in-one structure, where the NiFe alloy NPs on CeO2–NiFe@SCeFNM enhanced the electronic conductivity, and the CeO2 phase facilitated O2− migration as well as CO2 adsorption and activation, while the perovskite oxide framework ensured the structural integrity. The strong interaction between the three phases in the nanodomain creates a “win–win” effect, and results in an optimal CO2 electrolysis activity, which also presents a promising approach to enhancing SOEC performances by dual exsolution of A-site metal oxides and B-site metal NPs from perovskite oxides.

Author contributions

Biao Ouyang: conceptualization, investigation, writing – original draft preparation. Ming Yang: validation, data curation. Lin-Bo Liu: methodology. Shuo Liu: formal analysis. Yan Li: data curation. Xian-Zhu Fu: supervision. Yifei Sun: conceptualization. Subiao Liu: conceptualization, project administration, resources, supervision, writing – review and editing. Jing-Li Luo: supervision, validation.

Conflicts of interest

There are no conflicts of interest to declare.

Data availability

The data supporting this paper have been included as part of the supplementary information (SI). Supplementary information is available. See DOI: https://doi.org/10.1039/d5gc04390e.

Acknowledgements

This work is supported by the Pilot Group Program of the Research Fund for International Senior Scientists (Grant No. 22350710789) and the Start-up Funding of Central South University (No. 206030104). This project was also supported by Shenzhen Science and Technology Program (JCYJ20240813145707010, JCYJ20220530143401002, JCYJ20241202124410015). The authors are grateful for resources from the High-Performance Computing Center of Central South University.

References

  1. H. D. Matthews and S. Wynes, Science, 2022, 376, 1404–1409 CrossRef CAS PubMed.
  2. P. De Luna, C. Hahn, D. Higgins, S. A. Jaffer, T. F. Jaramillo and E. H. Sargent, Science, 2019, 364, eaav3506 CrossRef CAS PubMed.
  3. X. Hou, Y. Jiang, K. Wei, C. Jiang, T.-C. Jen, Y. Yao, X. Liu, J. Ma and J. T. S. Irvine, Chem. Rev., 2024, 124, 5119–5166 CrossRef CAS PubMed.
  4. J.-h. Myung, D. Neagu, D. N. Miller and J. T. S. Irvine, Nature, 2016, 537, 528–531 CrossRef CAS PubMed.
  5. T. L. Skafte, P. Blennow, J. Hjelm and C. Graves, J. Power Sources, 2018, 373, 54–60 CrossRef CAS.
  6. Y. Sun, Z. Liu, W. Zhang, X. Chu, Y. Cong, K. Huang and S. Feng, Small, 2019, 15, 1803513 CrossRef PubMed.
  7. J. Ni, L. Chen, J. Lin and S. Kawi, Nano Energy, 2012, 1, 674–686 CrossRef CAS.
  8. R. Xu, S. Liu, M. Yang, G. Yang, Z. Luo, R. Ran, W. Zhou and Z. Shao, Chem. Sci., 2024, 15, 11166–11187 RSC.
  9. M. Yang, L.-B. Liu, S. Liu, Y. Li, B. Ouyang, X.-Z. Fu, J.-L. Luo, Y. Sun and S. Liu, Chin. Chem. Lett., 2024, 110603 Search PubMed.
  10. Q. A. Islam, S. Paydar, N. Akbar, B. Zhu and Y. Wu, J. Power Sources, 2021, 492, 229626 CrossRef CAS.
  11. B.-W. Zhang, M.-N. Zhu, M.-R. Gao, X. Xi, N. Duan, Z. Chen, R.-F. Feng, H. Zeng and J.-L. Luo, Nat. Commun., 2022, 13, 4618 CrossRef CAS PubMed.
  12. H. Lv, T. Liu, X. Zhang, Y. Song, H. Matsumoto, N. Ta, C. Zeng, G. Wang and X. Bao, Angew. Chem., Int. Ed., 2020, 59, 15968–15973 CrossRef CAS PubMed.
  13. F. He, M. Hou, F. Zhu, D. Liu, H. Zhang, F. Yu, Y. Zhou, Y. Ding, M. Liu and Y. Chen, Adv. Energy Mater., 2022, 12, 2202175 CrossRef CAS.
  14. S. Jo, Y. H. Kim, H. Jeong, C.-h. Park, B.-R. Won, H. Jeon, K. T. Lee and J.-h. Myung, Appl. Energy, 2022, 323, 119615 CrossRef CAS.
  15. D. He, W. Ruan, J. Li, J. Ni and C. Ni, Chem. Eng. J., 2022, 433, 133632 CrossRef CAS.
  16. B.-W. Zhang, M.-N. Zhu, M.-R. Gao, J. Chen, X. Xi, J. Shen, R.-F. Feng, N. Semagina, N. Duan, H. Zeng and J.-L. Luo, Angew. Chem., Int. Ed., 2023, 62, e202305552 CrossRef CAS PubMed.
  17. B. Koo, K. Kim, J. K. Kim, H. Kwon, J. W. Han and W. Jung, Joule, 2018, 2, 1476–1499 CrossRef CAS.
  18. H. Téllez, J. Druce, J. A. Kilner and T. Ishihara, Faraday Discuss., 2015, 182, 145–157 RSC.
  19. Y. Tian, Y. Liu, A. Naden, L. Jia, M. Xu, W. Cui, B. Chi, J. Pu, J. T. S. Irvine and J. Li, J. Mater. Chem. A, 2020, 8, 14895–14899 RSC.
  20. J. Zhu, W. Zhang, Y. Li, W. Yue, G. Geng and B. Yu, Appl. Catal., B, 2020, 268, 118389 CrossRef CAS.
  21. E. Wang, C. Jin, L. Zhao, Z. Yang, C. Liu, S. Wang, X. Lei, M. Chao, H. Xu and R. Yang, Chem. Eng. J., 2024, 494, 153129 CrossRef CAS.
  22. Y. Yi, X. Xi, L. Huang, X. Liu, Y. Liao, J. Liu, J. Long, J. Zhang, X.-Z. Fu and J.-L. Luo, Adv. Funct. Mater., 2025, 35, 2412486 CrossRef CAS.
  23. S. Liu, H. Luo, Y. Li, Q. Liu and J.-L. Luo, Nano Energy, 2017, 40, 115–121 CrossRef CAS.
  24. L.-B. Liu, Y.-F. Tang, S. Liu, M. Yu, Y. Sun, X.-Z. Fu, J.-L. Luo and S. Liu, Adv. Energy Mater., 2025, 15, 2402967 CrossRef CAS.
  25. S. Liu, L.-B. Liu, M. Yang, Y. Li, Q. Pan, Y. Sun, X.-Z. Fu, S. Liu and J.-L. Luo, Chem. Eng. J., 2025, 519, 165404 CrossRef CAS.
  26. F. Hu, Y. Ling, S. Fang, L. Sui, H. Xiao, Y. Huang, S. Wang, B. He and L. Zhao, Appl. Catal., B, 2023, 337, 122968 CrossRef CAS.
  27. Y. Li, Y. Li, Y. Wan, Y. Xie, J. Zhu, H. Pan, X. Zheng and C. Xia, Adv. Energy Mater., 2019, 9, 1803156 CrossRef.
  28. Y. Luo, X. Chang, J. Wang, D. Zhang, L. Fu, X.-K. Gu, Y. Wang, T. Liu and M. Ding, ACS Nano, 2025, 19, 1463–1477 CrossRef CAS PubMed.
  29. Y. Song, X. Zhang, K. Xie, G. Wang and X. Bao, Adv. Mater., 2019, 31, 1902033 CrossRef CAS PubMed.
  30. L. Zhang, C. Xu, W. Sun, R. Ren, X. Yang, Y. Luo, J. Qiao, Z. Wang, S. Zhen and K. Sun, Sep. Purif. Technol., 2022, 298, 121475 CrossRef CAS.
  31. X. Xi, Y. Fan, J. Zhang, J.-L. Luo and X.-Z. Fu, J. Mater. Chem. A, 2022, 10, 2509–2518 RSC.
  32. L.-B. Liu, S. Liu, M. Yang, Y. Li, B. Ouyang, P. Zhang, X.-Z. Fu, Y. Sun, S. Liu and J.-L. Luo, Appl. Catal., B, 2025, 379, 125705 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.