Christian
Melcher
*a,
Andreas
Nenning
a,
Florian
Schrenk
b,
Kirsten
Rath
a,
Christoph
Rameshan
b and
Alexander Karl
Opitz
a
aTU Wien, Institute of Chemical Technologies and Analytics, Getreidemarkt 9/164-EC, 1060 Vienna, Austria. E-mail: chrisitan.melcher@tuwien.ac.at
bMontanuniversität Leoben, Chair of Physical Chemistry, Franz-Josef-Strasse 18, 8700 Leoben, Austria
First published on 11th March 2025
In solid oxide CO2 electrolysis cells, moderate activity and coking of the cathode are major issues that hinder commercialization of this important technology. It has been already shown that cathodes based on a mixed conducting oxide decorated with well-dispersed metal nanoparticles, which were grown via an exsolution process, are highly resilient to carbon deposition. Using perovskite-type oxides that contain reducible transition metals, such nanoparticles can be obtained in situ under sufficiently reducing conditions. However, the direct catalytic effect of exsolved metal nanoparticles on the CO2 splitting reaction has not yet been explored thoroughly (e.g. by employing well-defined model systems), thus, an in-depth understanding is still lacking. In this study, we aim at providing a crucial piece of insight into high-temperature electrochemical CO2 splitting on exsolution-decorated electrodes: we present the results of combined Near Ambient Pressure X-ray Photoelectron Spectroscopy (NAP-XPS) and electrochemical measurements on three different ferrite perovskites, which were employed as thin film model electrodes. The investigated materials are: La0.6Ca0.4FeO3−δ (LCF), Nd0.6Ca0.4FeO3−δ (NCF), and Pr0.6Ca0.4FeO3−δ (PCF). The results obtained allow us to directly link the electrode's CO2 splitting activity to their surface chemistry. Especially, the electro-catalytic activity of the materials decorated with and without metallic iron nanoparticles was in focus. Our experiments reveal that in contrast to their beneficial role in H2O electrolysis, exsolved Fe0 metal particles deteriorate CO2 electrolysis activity. This behavior contrasts with expectations derived from earlier reports on porous samples, and is likely a consequence of the differences between the CO2 splitting and H2O splitting mechanism.
Broader contextThe interplay between catalysis, materials science, and energy conversion is critical in addressing pressing global challenges such as climate change and carbon neutrality. High-temperature CO2 electrolysis is a promising approach to closing the carbon loop by converting CO2 into valuable chemical feedstocks or fuels. However, the role of surface chemistry and structural evolution in determining the performance and stability of catalytic materials remains poorly understood. In our work, we uncover the “dark side” of metal exsolution, which is known for enhancing catalytic activity for some reactions, by demonstrating its deactivating effects on CO2 splitting performance on perovskite-type oxides. Using well defined model-electrode systems, we show that CO2 splitting is fundamentally different from H2O splitting, as it relies on surface oxide reactions rather than metal-catalyzed recombination. Our findings emphasize the importance of tailored material design, particularly in controlling surface defect chemistry and lattice composition, to optimize catalytic performance for CO2 electrolysis. This study not only challenges conventional views on metal exsolution but also provides insights for developing advanced, coking-resistant materials for CO2 utilization. By addressing the mechanistic distinctions of CO2 and H2O splitting, we contribute to the development of efficient catalytic systems for sustainable energy applications. |
Well-established cathode materials for SOECs include Ni-YSZ cermets.1,4,5 However, these cathodes suffer from long-term degradation issues, including impurity poisoning, detrimental microstructure evolution,6,7 and a relatively high susceptibility towards carbon deposition, commonly referred to as coking.8,9 An alternative approach relies on using Ni-GDC (GDC = Gd2O3 doped CeO2) or pure oxides as cathode materials. In both cases, the tendency towards coking can be decreased drastically.10,11 However, especially for the latter, it is crucial that the material functions as a Mixed Ionic and Electronic Conductor (MIEC), since both the oxide ions as well as the electrons must be transported from/to the reaction site of CO2 reduction at the surface of the cathode material. This can be straightforwardly recognized by the respective cathodic half-cell reaction (which is given in eqn (1) in Kröger–Vink notation).
![]() | (1) |
Numerous studies have demonstrated that metallic phases on the surface of an oxide, such as exsolved nanoparticles, indeed enhance the rate of certain reactions. Examples include the reverse water–gas shift reaction (rWGS),34 methane dry reforming (MDR),20 or electrochemical water splitting.31,35,36 Crucial for all these reactions is to obtain an improved mechanistic understanding for the enhancement of reaction rates by exsolution particles, which in these cases always involves either enhanced adsorption or recombination rates of hydrogen species on the metallic particle surfaces. For instance, the proposed mechanism for water splitting and the reason for the beneficial effect of metallic particles on the reaction rate involves a spillover mechanism.36 The adsorbed hydrogen on the host oxide has a high surface mobility and can easily diffuse towards the metallic surfaces where recombination into H2 molecules is more favored than on an oxide surface.
In the case of CO2 electrolysis, however, it is highly questionable whether carbonate reaction intermediates offer fast enough surface diffusion rates to enable a similar spillover mechanism. Additionally, CO2 splitting does not require recombination of intermediates to produce CO. Moreover, there is a general lack of well-defined model studies on the effect of exsolution on electrochemical CO2 splitting. Studies claiming that exsolutions enhance CO2 splitting kinetics were typically performed on porous electrodes, on which also variations of morphology, ionic conductivity or oxide surface composition may be responsible for the observed performance differences. On the contrary, an in situ NAP-XPS study on well-defined model samples for direct CO2 electrolysis11 found no significant effect of metallic nanoparticles exsolved from perovskite-type ferrites and chromites on the CO2 splitting rate. Thus, the beneficial effect of metal exsolution for CO2 splitting remains questionable and is thus the main motivation behind this study.
From the viewpoint of solid state electrochemistry, the material La0.6Sr0.4FeO3−δ (LSF) serves as an ideal model electrode for exsolution studies, as its defect chemistry has been extensively studied and is thus well understood.31,37,38 Moreover, metallic iron particles can be exsolved from LSF by applying sufficiently strong cathodic bias.35,36,39 In the present work, we explore similar materials using the more abundant Ca as an A-site dopant instead of Sr, analyzing their electro-catalytic performance for direct electrochemical CO2 splitting. The materials under investigation are La0.6Ca0.4FeO3−δ (LCF), Nd0.6Ca0.4FeO3−δ (NCF), and Pr0.6Ca0.4FeO3−δ (PCF). The reason for substituting La with Nd and Pr (LCF → NCF, PCF) lies in the potential valence transitions (Pr3+ ⇌ Pr4+, Nd3+ ⇌ Nd2+) and the differences in ionic radii. Especially the latter might lead to differences in exsolution behavior and stability upon reduction.
As we operate under reducing conditions (pO2 ≈ 10−21 bar) the oxygen vacancy concentration is primarily determined by the extrinsic Ca dopant concentration. Corresponding Brouwer diagrams exist for very similar materials like LSF.40 According to these, the vacancy concentration is given by with the same dopant concentration for all three materials (40% Ca on the A-site). Upon biasing the electrodes, we change the electron concentration in our materials in the range of 0 to 10% Fe2+.40,41 Therefore, the oxygen vacancy concentration ranges from 0.2 to 0.25 per formula unit.
The objective of this paper is therefore to investigate whether exsolved Fe-nanoparticles indeed enhance the reaction rate of high temperature CO2 electrolysis. To answer this question conclusively, the experimental setting was highly optimized and the experimental boundary conditions were tailored even better to this goal:
(i) Rather than using pure CO2 (which was needed for determination of CO2 conversion rates in ref. 11), we went for a CO:
CO2 mixture of 1
:
10 to obtain a thermodynamically well-defined p(O2) for all our measurements leading to a well-defined reference point for the overpotential.
(ii) Materials wise, we focused purely on Ca-doped ferrites (i.e. all studied perovskite-type materials have only Fe on the B-site) with the A-site being either La, Nd or Pr (LCF, NCF or PCF). Such a focused approach is crucial to gain better mechanistic understanding of the interplay between A-site composition, Fe-exsolution and the CO2 splitting reaction rates. This is achieved by utilizing a lab-based Near Ambient Pressure X-Ray Spectrometer (NAP-XPS) in tandem with Electrochemical Impedance Spectroscopy (EIS) and Direct Current (DC) measurements. Through this technique, the electrochemical performance and the surface composition can be studied simultaneously.
(iii) Especially, we compare the reaction rates before and after Fe-exsolution under otherwise identical conditions (same overpotential, same p(O2), same temperature, same surface area, same sample) in one experiment.
(iv) In order to rule out as many additional contributions to the surface activity as possible, geometrically well-defined dense model electrodes were fabricated using Pulsed Laser Deposition (PLD) and photolithography. This ensures that the model electrodes have a well-defined surface area, enabling comparisons across different materials and linking electrochemical activity directly to the surface chemistry.
Before the actual working electrode (WE) was deposited, a 15|5 μm (mesh|strip width) Ti/Pt current collector (thickness of 5 nm Ti below 100 nm Pt) was prepared on the polished side of the YSZ single crystal electrolyte by magnetron sputtering (BalTec MED 020, Leica Microsystem GmbH, Germany) and photolithography. For a more detailed description of the photolithography process, please refer to ref. 39. Subsequently, an about 100 nm thin layer of the desired perovskite-type material (LCF, NCF or PCF) was grown by PLD atop the current collector at a pO2 of 4.0 × 10−2 mbar, using a KrF excimer laser (1800 pulses, 5 Hz, λ = 248 nm, COMPex Pro 201, Lambda Physics).43 This laser was employed to ablate the perovskite-type target (LCF, NCF, or PCF) that had been prepared via a modified Pechini method using citric acid as complexing agent,44 followed by isostatic pressing of the obtained powder and sintering in air at 1150 °C for 12 h. During thin film deposition, the temperature of the YSZ substrate was set to 600 °C (achieved by resistive heating and temperature control using a pyrometer).
The spectrometer utilized was a lab-based configuration equipped with a monochromatic Al-Kα X-ray source (μFOCUS 500 NAP, SPECS, Germany). In this setup, the hemispherical analyzer (PHOIBOS 150, SPECS, Germany) is positioned behind differentially pumped electrostatic lenses. To minimize gas phase scattering, the sample surface was positioned very close to the water-cooled nozzle, approximately 500 μm away. For a more detailed explanation of the NAP-XPS setup, please refer to ref. 46.
The sample holder, specifically designed for such experiments by Huber Scientific, consists primarily of corundum (Al2O3). The sample was fastened using four Pt pins, which were pressed against the corners of the sample through a screwing mechanism. While these pins provided electrical contact to the WE, the CE was separately connected using a Pt wire. An aperture on the backside of the sample holder permits an IR laser beam to irradiate and thus heat up the sample. The resulting temperature was monitored using a pyrometer. For an accurate temperature control, the emissivity of each individual sample was calibrated using EIS measurements as follows: the temperature dependence of the ionic conductivity of YSZ is well known.47 Since the geometry of the YSZ single crystal is known as well, the ohmic resistance of the sample (RYSZ ≈ x-axis intercept in the Nyquist-plot) can be used to calculate the sample's temperature. This works best at relatively low temperatures (300 °C to 500 °C), since contributions of wiring and contact resistances to the total ohmic resistance can be neglected. At these lower temperatures the emissivity used by the pyrometer was adjusted to match the calculated temperature from EIS measurements. When going to higher temperatures (600 °C to 800 °C), the now calibrated pyrometer was then used for temperature control. The laser's power output was regulated through pulse width modulation according to ref. 46.
As a cleaning routine, all samples were initially heated up to 600 °C in 1 mbar O2 (99.999% purity, Messer, Germany) for one hour inside the NAP-XPS chamber to remove carbon contaminants before the start of the CO2 splitting experiment. For the subsequent NAP-XPS experiments, a total gas pressure of 1 mbar was established with a CO:
CO2 ratio of 1
:
10 (CO2: 99.9995% purity, Messer, Germany; CO: 99.994% purity, Air Liquide, France) ensuring a constant and well-defined effective pO2 in the gas phase for each temperature, which determines the oxygen non-stoichiometry of our oxide materials.
Electrochemical measurements were facilitated inside the NAP-XPS chamber by feedthrough connections via a CF 63 flange designed for BNC connectors. For DC and EIS measurements, a Source Meter Unit (SMU, 2410 SourceMeter, Keithley Instruments, United States) and an impedance analyzer (Alpha-20 A High Performance Frequency Analyzer, Novocontrol Technologies GmbH & Co. KG, Germany) were employed, respectively. For electrochemical impedance spectroscopy (EIS), an AC voltage of 10 mV (root mean square) was applied in addition to the constant DC bias. For all experiments, the WE was connected to the mass of the XPS analyzer, and electrochemical polarization in potentiostatic mode was achieved by applying the inverse polarization to the CE. For I–V curve measurements, time-resolved potentiostatic DC measurements were conducted using the SMU.
For recording current–voltage characteristics (I–V curves), a bias voltage was applied to the model cell, which leads to cathodic polarization of the WE. This cathodic polarization was increased stepwise up to the point where Fe-exsolution was clearly visible in the Fe 2p region of the NAP-XPS spectra (roughly −275 mV, depending on the studied material). Subsequently, the cathodic polarization was decreased (i.e. to more positive overpotentials) to correlate the amount of metallic Fe on the surface of the electrodes to the current density. A schematic diagram of the applied voltage over the course of an experiment is shown in Fig. 1c. Time-resolved IDC values were recorded after the bias was increased to distinguish capacitive current from faradaic current. After the IDC values reached a steady state, EIS measurements were performed under the given bias.
As both electrodes were located in the same atmosphere, the oxygen activity in the WE a(O2,WE) could be calculated according to Nernst's equation given by eqn (2), where η and Uvs.CE correspond to the overpotential dropping at the WE and the applied voltage Uvs.CE with respect to the CE, respectively. R, T and F have their usual meanings.
![]() | (2) |
In principle, the overpotential η differs from the applied bias because the electrolyte, the CE, the wiring as well as the contact resistances contribute to the overall voltage drop (η = Uvs.CE − IDC(RYSZ + RCE + Rwire + Rcontact)). However, considering the high resistances of our thin-film electrodes, especially at the low gas pressure of 1 mbar, the former mentioned contributions can safely be neglected (i.e. η ≈ Uvs.CE holds in good approximation). The effective overpotential dropping at the WE is therefore close to the applied voltage with deviations of at most 4% (worst case at maximal applied overpotential and maximal current at 800 °C).
![]() | (3) |
For the evaluation of the area specific resistance (ASR) and the chemical capacitance (Cchem) values, a complex non-linear least squares fitting function is calculated using a RYSZ–Rode‖CPE equivalent circuit, which allows extracting the electrode polarization resistance Rode (the diameter of the semicircle in the Nyquist-plot) and the capacitance C from the low frequency feature. For the ASR, Rode is normalized to the geometric surface area (roughly 0.2 cm2).
The temperature dependence of the electrode activity is displayed in an Arrhenius plot (Fig. 2b) where the reciprocal values of the area specific resistances of Rode (ASR−1) are plotted against 1000/T (again solid markers indicate measurements at OCV, hollow markers represent values under cathodic η of −200 mV vs. CE). The activation energies Ea, which can be calculated from the slopes of linear fits of the data points in Fig. 2b (solid lines), range between 0.56 ± 0.03 eV and 0.81 ± 0.05 eV (uncertainties represent two times the standard deviation derived from the uncertainties in the slopes of the least-squares fit in the Arrhenius plot, corresponding roughly to a 95% confidence interval). These values remain largely unaffected when a cathodic overpotential is applied, however, the ASR significantly decreases under bias causing a parallel shift in the Arrhenius plots. This behavior is expected and aligns with the understanding that perovskite-type mixed conducting electrodes in reducing atmospheres typically exhibit non-linear I–V curves, which was already observed for LSF electrodes in ref. 11 and 37.
The capacitance associated with the low-frequency feature of the impedance spectra, captured in the fit as a constant phase element (CPE), represents the largest capacitance in the system. Based on our defect chemical understanding of closely related perovskites such as LSF38 it can be safely assumed that this capacitance corresponds to the chemical capacitance Cchem of the perovskite-type working electrodes rather than to other typically electrostatic capacitive contributions (e.g. originating from interfaces).
As Cchem depends on the minority charge carrier concentration38 and thus scales with the film thickness, the capacitances (in the order of 10−4–10−3 F) are normalized to the volume of the film (roughly 2 × 10−6 cm2), which leads to values in the order of 102–103 F cm−3 at 700 °C (see Fig. 2c). The film thickness was determined using SEM cross sections (see Fig. S11–S13 in the ESI†) and vary between 240 and 360 nm. Error bars in Fig. 2c depict the uncertainties of the Cchem values which are propagated by two times the standard deviation from SEM cross section thicknesses (roughly 10%). Uncertainties coming from impedance fitting are neglected, as they are found to be much smaller than the scatter in film thicknesses (roughly 0.2%).
As an RYSZ–Rode‖CPE circuit was used for impedance fitting, with the CPE element covering the slight deviations from an ideal semicircle,52 a real capacitance can be calculated in good approximation from the CPE-element fit parameters and the electrode resistance according to ref. 53. The obtained Cchem is directly proportional to the minority charge carrier concentration of the WE bulk material.54,55 Furthermore, the defect chemistry of the given ferrite-based perovskites in reducing atmospheres is governed by a high number of oxygen vacancies, and a comparatively small concentration of electronic point defects, which are expected to be localized to the B-site cations of the perovskite, thus appearing as Fe2+ ions.41 Under these conditions, and assuming that defect interactions play a minor role, the Fe2+ bulk concentration [Fe2+] can be calculated according to eqn (4).38 Here, Vm and Vfilm represent the molar volume and the volume of the PLD layer, respectively, while R, T and F retain their usual meanings.
![]() | (4) |
In order to comprehensively assess the surface chemistry of the electrodes under in situ conditions, it is not only important to spectroscopically analyze the constituents of the electrode materials, but also those of potential adsorbates. From literature it is known that carbonate-type intermediates form on the electrode surface during CO2 electrolysis on ceria-based9,56 and perovskite-type electrodes.11 Therefore, the observation of these intermediates under conditions without Fe metal exsolution is a first important step.
NAP-XPS spectra of the C 1s and O 1s regions for LCF under an overpotential of −200 mV vs. CE across varying temperatures are presented in Fig. 4a. For better comparability, the spectra in the C 1s region are normalized to the baseline of each individual C 1s spectrum. Three main peaks are evident: a CO2 gas phase peak at 292.8 eV,11,57 a peak related to a carbonate adsorbate at 289.5 eV,11,57,58 and – for the lowest temperature – an asymmetric peak at 283.8 eV corresponding to graphite-like carbon.11,57,59–61 The O 1s spectra in Fig. 4a display a pronounced asymmetric peak at 528.8 eV, stemming from lattice oxygen species and in part from surface oxygen species.57,62,63 This peak is modeled using two components: Olattice and Oasym. Additionally, a signal at 531.4 eV emerges, corresponding to an adsorbed carbonate-type species.11,57Fig. 4b displays the normalized intensities of the O 1s regions for all three materials at OCV and under −200 mV overpotential. The Ocarbonate peak is evident regardless of the material and increases with cathodic overpotential.
The graphite-like carbon peak for LCF at −200 mV in the C 1s region in Fig. 4a is only visible at the lowest temperature of 500 °C. Signals from the carbonate species in both the C 1s and O 1s regions decrease above 600 °C. For further quantification of the carbonate coverage, the O 1s region is considered, as it offers a superior signal-to-noise ratio (SNR) in comparison to the C 1s spectra. However, to observe graphite-like carbon on the surface, the C 1s region is essential.
Analogous NAP-XPS spectra were acquired for all three electrode materials under various overpotentials. Fig. 5a displays heatmaps of the Ocarbonate/Otot intensity ratios as a function of temperature and overpotential. Each square represents a set of temperature and overpotential and the color of each square indicates the intensity ratio. The adsorbed carbonate is observed under all conditions and the intensity ratios range from 2% to 25%. The signal intensity ratio Cgraphite/Ctot in Fig. 5b on the other hand is zero under most conditions and graphite-like carbon is only observed at lower temperatures (500 °C and 600 °C) and higher cathodic overpotentials.
Fig. 6a shows the Fe 2p region of the measurements on LCF at 700 °C under varying cathodic overpotentials (η). Note that the XPS fitting is not sufficient to separate Fe2+ from Fe3+. Therefore, only the sum of both is depicted as Fe oxide which is modeled using two constrained peaks to match the asymmetric shape of the Fe oxide peak. The Fe metal peak on the other hand is clearly separated and well de-convoluted using XPS fitting.
Initially, η was set from −200 mV to −300 mV in a stepwise manner (increasing |η|, spectra no. 1–3 in Fig. 6a). Subsequently, η was retraced back to −250 mV (decreasing |η|, spectrum no. 4 in Fig. 6a). It is evident that initially, under overpotentials slightly less cathodic than or equal to −250 mV, no metallic iron is present on the surface, even though we would expect it to be stable under these conditions. Only under overpotentials more cathodic than −250 mV, two additional peaks appear at 707 eV and 719 eV for the Fe 2p3/2 and Fe 2p1/2 branches. Their binding energies perfectly match metallic Fe,31,64,65 indicating the successful exsolution of Fe0 particles. These additional peaks remain unaffected when η is retraced back from a maximum −300 mV to −250 mV (decreasing |η|, spectrum no. 4 in Fig. 6a) and the exsolved iron remains metallic even under conditions at which no Fe0 was observed during the initial reduction (compare spectra no. 2 and 4 in Fig. 6a). This demonstrates that the initial formation of exsolved particles requires an additional exsolution overpotential ηex, which is in line with our prior experience with similar materials as well as with literature reports. Consequently, the experimental strategy proposed in Fig. 1c is directly applicable to the materials studied here. This offers the opportunity that two different states of the material can be observed under otherwise identical conditions (same overpotential and temperature): one with and one without Fe exsolution. In other words, a hysteresis-like behavior emerges regarding the Fe0 content on the surface. As a next step, the surface activities of these different stages (with and without exsolved Fe0) will be compared in the following chapter by looking at the current voltage characteristics (I–V curve).
The tendency for Fe exsolution of different materials can be an important factor for the kinetics of CO2 splitting. Therefore, a comprehensive visual representation of the relative Fe0 amount is illustrated using heatmaps in Fig. 6b for the increasing |η| branch thus depicting the Fe exsolution tendency of the pristine materials. Here, the NAP-XPS signal intensity ratios of Fe0/Fetot (Fetot = Fe0 + FeII + FeIII) from the Fe 2p region are displayed as a function of temperature and overpotential for all three materials. Caused by uncertainties from XPS fitting, the Fe0/Fetot threshold indicating significant amounts of exsolution was found to be 3%. Therefore, Fe0/Fetot ratios below this 3% threshold are assumed to be insignificant, which is taken into account in the definition of the color bar in Fig. 6b. The results show that the amount of metallic Fe is strongly increased by overpotential. Temperature on the other hand seems to play only a minor role.
Beneath the I–V curve of each electrode in Fig. 7 the fraction of metallic Fe0 on the surface of the electrodes (Fe0/Fetot as calculated from NAP-XPS measurements) is plotted also as a function of the applied overpotential. As was already described in Fig. 6a, all materials show a hysteresis-like relationship with respect to the Fe0 fraction as well, and the overpotential region of the Fe0 fraction hysteresis loop perfectly coincides with the I–V hysteresis. Similar I–V and Fe0 fraction curves for all three materials across the entire measured temperature range (500 °C to 800 °C) can be found in Fig. S4 in the ESI.† The results show that the above-described hysteresis in Fe0 fraction and the electrode de-activation are reproducible for all materials and temperatures.
The electrochemical impedance (Fig. 2 and Fig. S2, S3, ESI†) as well as DC measurements (Fig. 7 and Fig. S4, ESI†) show a superior electrochemical activity of LCF, compared to NCF and PCF, across the entire measured temperature range. This difference in electrochemical activity of the three perovskite-type electrodes, while exhibiting virtually identical A:
B cation ratio at the surface, may be explained by two effects: either, the rare earth metal on the A-site also contributes redox activity, or the different size of La, Nd, and Pr affects the defect chemistry of the perovskite.
By looking at the NAP-XPS spectra of the A-site elements (La 3d, Nd 3d and Pr 3d) in Fig. 3a it becomes evident that no changes regarding the shapes of the spectra are observed upon electrochemical reduction. Only a binding energy shift of the entire A-site spectrum is observed, which can be explained by the Fermi-level freely moving through the band gap without being pinned by a redox active state.39 Since the shapes of the spectra are not altered at all after the reduction, a redox change can be ruled out for all three A-site elements. Furthermore, since La can be expected to be the least redox active element of the three, but LCF offers the fastest kinetics, the first option – i.e. the rare earth element contributing redox activity – can be excluded as an explanation for the materials’ different electrochemical activity.
Hence, the size effect of the A-site cations may be the more suitable explanation. Indeed, the observed activity trend (LCF > PCF ≈ NCF) correlates with the ionic radii of the elements: La > Pr ≈ Nd. The larger ionic radius of La3+ leads to a lattice expansion, as evidenced by XRD, which reveals that LCF exhibits the largest unit cell (see Fig. S1 and Table S1 in ESI†). This lattice expansion can increase the reducibility of the perovskite and enhance its surface reactivity, a phenomenon previously demonstrated for similar perovskites through 18O tracer exchange experiments on intentionally tensile-strained cobaltite perovskite thin films.71 Similar A-site size effects on the reducibility of perovskite-type ferrites were reported using thermogravimetry.72 Also, for the materials studied in the present case, an easier reducibility of LCF is supported by EIS measurements, since LCF indeed exhibits the largest chemical capacitance Cchem. Since Cchem is directly proportional to the minority charge carrier, this indicates a larger n-type electron concentration (i.e. Fe2+ concentration)54 and thus a higher degree of reduction of LCF than the other two oxides under otherwise identical conditions (compare Fig. 2c and Fig. S3 in the ESI†). From the viewpoint of LCF exhibiting a larger lattice constant this behavior seems to be plausible, as a lattice with larger unit cell volume can easier accommodate electronic charge carriers that localize at the B-site cation thus appearing as Fe2+, which is larger than Fe3+. Interestingly, Fe metal exsolution (which does not necessarily correlate to Fe3+/Fe2+ reduction) on the other hand is more pronounced for NCF and PCF. This phenomenon might be attributed to different types/degrees of lattice strain caused by different sizes of the A-site cations. Caused by the relatively smaller Nd3+ and Pr3+ cations compared to La3+, NCF and PCF exhibit more lattice distortion (and thus more local strain) compared to LCF. Strain has already been demonstrated to be able to strongly affect the exsolution behavior of perovskite films,73 and hence misfit-strain induced by size-mismatch of the A-site element may be a possible explanation for the observed differences of LCF, NCF, and PCF.
Regarding the effective pO2 of the initial formation of exsolution particles, the measurements are in line with literature. Very similar exsolution onset points of Fe0 nanoparticles were previously observed for NCF under steam electrolysis conditions in ref. 74. The Fe0/FeO equilibrium at 700 °C corresponds to an equivalent pO2 of 2.7 × 10−22 bar,75 which corresponds to a Nernst voltage of −114 mV vs. CE in a 1:
10 CO
:
CO2 atmosphere (pO2 of 6.4 × 10−20 bar). The fact that we observe the Fe0 exsolution onset at substantially more cathodic overpotentials (about −275 mV) strongly suggests that the initial exsolution process is kinetically hampered – e.g. by diffusion of involved species, electron transfer of the required iron reduction reaction, particle nucleation, and particle growth – and is in line with previous reports.74,76
When comparing the exsolution onsets of the different materials, subtle differences emerge by looking at Fig. 6b. Firstly, LCF appears to exhibit a slightly lower tendency for Fe exsolution since the Fe0 fraction at high overpotentials is rather low. Secondly, while LCF seems to exsolve in a more gradual manner, both NCF and PCF display a more abrupt increase in the Fe0 fraction upon reaching a certain overpotential threshold at around −275 mV.
Analyzing the iDC values of both states in Fig. 7, a clearly negative correlation between iDC and Fe0 fraction emerges, since the magnitude of the current is decreasing with increasing Fe0 fraction. The onsets of the hysteresis loops of both the iDC values and Fe0 fractions also coincide almost perfectly. Furthermore, the overpotential for the exsolution onset is within the range where the iDC values start to deviate from an exponential behavior for LCF and PCF. Since gas diffusion limitation can be safely excluded for thin films in an atmosphere of 1 mbar,77 this behavior suggests the onset of a surface de-activation process. This clearly suggests that Fe exsolution appears to affect the reaction rate, but in a detrimental manner.
The negative correlation between |iDC| and the Fe0 fraction is reproducibly demonstrated for all three investigated materials across a wide temperature range. Analogous results at different temperatures are provided in the ESI† (Fig. S4). The consistent manifestation of this phenomenon is noteworthy: a hysteresis-like behavior for both the current density and the Fe0 amount is evident across almost all material and temperature combinations, along with the negative correlation between them.
Furthermore, quantification using electrochemical (bulk) data revealed that for PCF (which exhibits by far the highest Fe0 fraction) at 700 °C, roughly 23% of the total Fe atoms present in the lattice were reduced to metal (see Table S3 in ESI†). Given the magnitude of these B-site vacancy concentrations, it seems improbable for the perovskite lattice to remain stable.32,41 De-activation caused by partial decomposition of the perovskite host would indeed be a plausible explanation at first glance. However, this does not necessarily seem to be the case here. Remarkably, once the exsolved Fe0 is re-oxidized, the surface activity is mostly recovered. This observation appears surprising, and explanations for this phenomenon remain speculative. Owing to the complexity of the observed effect, a detailed explanation is beyond the scope of this paper. Nevertheless, some thoughts deserve attention at this point: on the one hand, (at least partial) reversible exsolution could be an explanation for the re-activation upon re-oxidation. However, from what is known from literature, this seems unlikely given that exsolution – at least at the rather moderate temperatures we applied here – is typically considered to a great extent irreversible.32,36 This also suggests that the decrease in activity upon exsolution is not a sheer effect of losing surface area due to coverage by the exsolved metal particles, since the oxidized particles still remain after re-oxidation. Possibly, the oxidation state of the surface-decorating particles indirectly affects the reactivity of the perovskite surface, e.g. via triggering work function changes of the oxide. Another possibility would be a change in morphology of the particle after oxidation caused by a sudden change in the nanoparticle/perovskite interface energy. In addition, the evolution of surface roughness upon exsolution and re-oxidation may also affect the net activity of the electrode. SEM measurements revealed pronounced morphological changes of the thin-film electrodes when comparing post-measurement samples with a pristine one (see Fig. S7–S10 in the ESI†). However, the recovery of electrode surface activity upon particle re-oxidation can only hardly be explained to the full extent from the data available so far, and additional experiments are required to draw clear conclusions.
The de-activation of the electrodes for CO2 splitting upon Fe0 exsolution seems surprising at first glance, since the beneficial effect of exsolution was demonstrated for many reactions e.g. steam electrolysis.36,74 However, the reaction mechanisms for CO2 and H2O splitting differ significantly. In H2O splitting, the metal particles primarily facilitate the recombination of two neutrally adsorbed hydrogen atoms Had to form H2(36). In contrast, for CO2 splitting the presence of a metal appears not to enhance the desorption process of CO. Instead, the entire reaction seems to proceed entirely on the oxide surface. Its rate-limiting step is likely the conversion of a carbonate-type intermediate, which involves an electron transfer from the oxide to the carbonate.11 This carbonate intermediate was also observed in the present study and will be discussed in detail in the following section.
![]() | (5) |
![]() | (6) |
To put the NAP-XPS intensity ratios Ocarbonate/Otot into perspective, the carbonate surface coverage θcarbonate is roughly approximated by a simple calculation. With an inelastic mean free path of the O 1s photoelectrons of 2 nm78 and a bulk oxygen density of 5 × 1022 oxygen atoms cm−3, the bulk O 1s signal stems from approximately 1016 oxygen atoms cm−2. When further considering the C:
O ratio of 1
:
3 of the carbonate, its surface concentration can be estimated by eqn (7). The maximum carbonate intensity in Fig. 5a corresponds to a coverage of about 63% relative to the unit cell density of [100] oriented film (3 × 1015 unit cells cm−2). Therefore, we can assume that already more than half of a monolayer is present at 600 °C and high overpotential. While this approximation is subject to uncertainties, it still indicates a relatively high carbonate coverage, supporting the idea that carbonate conversion is the rate-limiting step.
![]() | (7) |
![]() | (8) |
Coking was primarily observed at 500 °C and high overpotentials (Fig. 4 and 5), which might further affect the carbonate coverage on the oxide surface since some adsorption sites might be blocked. At higher temperatures graphite-like carbon does not form, confirming the coking resistance of these materials. This is consistent with the findings of Skafte et al.,9 who observed that high oxygen vacancy concentrations and carbonate coverages on the surface delay graphite-like carbon formation.
![]() | ||
Fig. 8 Sketch for comparing the key pathways for (a) H2O and (b) CO2 splitting. Solid black arrows depict the fast/dominant processes, dashed grey arrows depict slow/unlikely processes. |
For CO2 splitting in Fig. 8b, the reaction proceeds differently. First, the carbonate intermediate – especially in the case of a carbonate bidentate11 – is probably significantly less mobile on the oxides surface and therefore might not be able to diffuse towards the exsolved nanoparticle with significant rate. Second, for CO formation, no recombination of adsorbed species is required. Therefore, the reaction on the oxide surface is probably still the dominant route for CO formation. Thus, no beneficial effect of metallic nanoparticles is observed.
• LCF displayed superior surface activity and the highest chemical capacitance. For all A-site elements, no changes in valence states were observed using NAP-XPS, confirming that the A-site predominantly acts as a structural provider for the materials. The key aspect for LCF's superior activity may therefore be the higher ionic radius of La3+ compared to Nd3+ and Pr3+. With larger A-site cations the perovskite lattice expands, thus being able to accommodate higher concentrations of n-type electronic charge carriers (i.e. higher concentrations of the larger Fe2+). Consequently, LCF exhibits more Fe2+, while NCF and PCF tend to exsolve higher amounts of metallic Fe0 under the same conditions.
• While Fe exsolution had been proven beneficial for high-temperature H2O splitting, it consistently exhibits a de-activating effect on direct electrochemical CO2 splitting in our study. This observation is validated by the tandem hysteresis behavior observed in current density and Fe0 fraction measured by in situ NAP-XPS. A plausible explanation for this substantial difference between H2O and CO2 splitting lies in the fundamental differences in their reaction mechanisms. For H2O splitting, the recombination of neutrally adsorbed hydrogen on the surface is accelerated strongly by metal particles via a spillover mechanism. In contrast, for CO formation in CO2 splitting, no such recombination is required and the reaction occurs on the oxide surface rather than on the metal. Additionally, the surface diffusivity of carbonate intermediates is probably quite low, preventing them from migrating towards the metallic nanoparticles.
• Using NAP-XPS, the well-known carbonate-type intermediate was detected. Notably, all three studied perovskite-type oxide electrodes showed excellent coking resistance, as graphite-like carbon was only identified in the lower temperature range of 500 °C to 600 °C. At higher temperatures, no graphite-like carbon was observed, demonstrating their potential as low-degrading electrode materials.
In conclusion, our research underscores the critical role of fundamental studies on well-defined model samples, which are essential for understanding reaction dynamics and the interplay between chemical surface states and electrochemical activity. Moreover, it is crucial to recognize that correlations observed for specific reactions may not apply universally, as demonstrated by the differing impact of Fe exsolution on H2O splitting versus direct CO2 splitting.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ey00013k |
This journal is © The Royal Society of Chemistry 2025 |