Next Article in Journal
A Data-Driven Identification Procedure for HVAC Processes with Laboratory and Real-World Validation
Previous Article in Journal
Recovery of Household Waste by Generation of Biogas as Energy and Compost as Bio-Fertilizer—A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impact of Silica Additions on the Phase Composition and Electrical Transport Properties of Ruddlesden-Popper La2NiO4+δ Mixed Conducting Ceramics

by
Kiryl Zakharchuk
1,
Aleksandr Bamburov
1,
Eugene N. Naumovich
2,3,
Miguel A. Vieira
1 and
Aleksey A. Yaremchenko
1,*
1
CICECO—Aveiro Institute of Materials, Department of Materials and Ceramic Engineering, University of Aveiro, 3810-193 Aveiro, Portugal
2
Department of High Temperature Electrochemical Processes, Institute of Power Engineering, Mory 8, 01-330 Warsaw, Poland
3
CTH2—Center for Hydrogen Technologies, Institute of Power Engineering, Augustowka 36, 02-981 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Processes 2022, 10(1), 82; https://doi.org/10.3390/pr10010082
Submission received: 2 December 2021 / Revised: 21 December 2021 / Accepted: 28 December 2021 / Published: 31 December 2021
(This article belongs to the Section Energy Systems)

Abstract

:
The present work explores the possibility of incorporation of silicon into the crystal structure of Ruddlesden-Popper La2NiO4+δ mixed conducting ceramics with the aim to improve the chemical compatibility with lanthanum silicate-based solid electrolytes. Ceramics with the nominal composition La2Ni1−ySiyO4+δ (y = 0, 0.02 and 0.05) were prepared by the glycine nitrate combustion technique and sintered at 1450 °C. While minor changes in the lattice parameters of the tetragonal K2NiF4-type lattice may suggest incorporation of a small fraction of Si into the Ni sublattice, combined XRD and SEM/EDS studies indicate that this fraction is very limited (≪2 at.%, if any). Instead, additions of silica result in segregation of apatite-type La10−xSi6O26+δ and La2O3 secondary phases as confirmed experimentally and supported by the static lattice simulations. Both total electrical conductivity and oxygen-ionic transport in La2NiO4+δ ceramics are suppressed by silica additions. The preferential reactivity of silica with lanthanum oxide opens a possibility to improve the compatibility between lanthanum silicate-based solid electrolytes and La2NiO4+δ-based electrodes by appropriate surface modifications. The promising potential of this approach is supported by preliminary tests of electrodes infiltrated with lanthanum oxide.

1. Introduction

Solid oxide fuel cells (SOFCs) are one of the most promising technologies for future power generation with expected gains in efficiency compared to conventional combustion-based technologies and with a wide range of potential applications including power sources for remote areas, transportation, and small to large-scale stationary power applications [1,2,3]. In addition to high energy conversion efficiency, the advantages of SOFC technology include fuel flexibility (from hydrogen to natural gas and biogas), low emissions and noiseless operation. The research in the SOFC area in recent years has been focused on lowering the operating temperature to 500–800 °C, thus aiming to reduce the system cost and improve cell durability [3].
Rare-earth silicates with apatite-type structure have attracted significant attention in the last two decades as a promising group of comparatively low-cost materials for use as solid electrolytes in the intermediate-temperature SOFCs [4,5,6,7,8,9]. The compounds have the general formula M10(SiO4)6O2+z, where M is a rare-earth or alkaline-earth metal. Lanthanum silicates exhibit the highest conductivities and show nearly pure oxygen-ionic conductivities over a wide range of oxygen partial pressures [10,11,12,13,14]. A number of studies of chemical substitutions aimed at optimizing the ionic conductivity showed that silicate apatites can tolerate a wide range of dopants and that the conductivity depends on the degree of anion or cation nonstoichiometry. The level of oxygen-ionic transport in lanthanum silicates with optimized composition is higher than in state-of-the-art yttria-stabilized zirconia electrolytes below 800 °C and achieves values > 0.001 S/cm at 500 °C being comparable with the highest conductivity reported for other solid electrolytes such as doped ceria and lanthanum gallate [8,9,10,11,12].
Despite substantial interest in apatite-type solid electrolytes, there have been a rather limited number of reports on the behavior of electrode materials in contact with these silicates [7,15,16,17,18,19,20,21,22,23,24,25,26,27]. Analysis of the available literature indicates that it is the chemical incompatibility between cell components and insufficient electrochemical performance of electrodes in contact with silicate-based solid electrolytes that may hinder their implementation in the SOFC technology. Typically, no formation of insulating phases between the lanthanum silicate-based electrolyte and the electrode materials was detected even at temperatures above 1100 °C [7,19,20,21,22,23,24,25]. However, noticeable cation interdiffusion between the electrolyte and electrode at the interface and silica diffusion from the electrolyte surface is often observed [7,15,16,22,23,24,25]. These factors may result in the deterioration of transport properties at the interfacial layers of both electrode and electrolyte and partial blocking of the electrochemical reaction zone, and, therefore, seem to be responsible for the rather poor behavior of tested cathode and anode materials. This is verified by the observation that polarization resistance increases with electrode fabrication temperature [7,24]. Among tested cathode materials, perovskite-type manganites showed the worst performance [7,15,16,21,22,23] due to significant cation interdiffusion and low ionic conductivity at the intermediate temperatures. Although attractive results were reported for perovskite-related Ruddlesden-Popper Ln2NiO4+δ-based nickelates, overpotential values are still high at temperatures below 800 °C [19,22,23,26].
Rare-earth nickelates Ln2NiO4+δ belong to the family of layered Ruddlesden-Popper (RP) phases with the general formula An+1BnO3n+1. The RP-type structure of Ln2NiO4 with n = 1, or K2NiF4-type structure, is built of perovskite-type LnNiO3 layers alternating with rock-salt-type LnO layers along the crystallographic c axis. A characteristic feature of Ln2NiO4+δ nickelates is their ability to accommodate excess oxygen that is incorporated in the interstitial positions in the rock-salt-type layers and is responsible for oxygen-ionic transport in these phases [28,29,30,31]. Substantial electronic conductivity (~50–100 S/cm at 500–800 °C) and high oxygen diffusion coefficients ensure the competitive electrochemical performance of mixed-conducting Ln2NiO4+δ-based oxygen electrodes in SOFCs [28,29,30,31,32].
As the Ln2NiO4 structure is very flexible to different kinds of substitutions in both sublattice, one possible approach to improve the compatibility of Ln2NiO4-based electrodes with apatite-type lanthanum silicate solid electrolytes may be the incorporation of silicon cations into the Ln2NiO4 lattice aiming to reduce the chemical potential gradient of silicon and therefore the driving force for unfavorable silica diffusion. The substitutions by silicon into perovskite-related oxides for electrochemical applications are not common but generally can be implemented [33]. In particular, the possibility of partial substitution of B-site transition metal cations by silicon and the formation of solid solutions were reported for SrM1−ySiyO3−δ (M = Co, Fe, Mn) [34,35,36], Sr1−xCaxM1−ySiyO3−δ (M = Mn, Fe) [37,38], A2MnFe1−ySiyO6−δ (A = Ca, Sr) [39], La0.6Sr0.4Co1−yzFeySizO3−δ [40] and Sr0.9Y0.1Co1−ySiyO3−δ [40] systems, often with positive effects on structural properties and electrical conductivity.
Thus, the goal of this work was to explore the possibility of incorporation of silicon cations into the nickel sublattice of La2NiO4+δ and to evaluate the impact of silica additions on the electrical transport properties. The phase formation was analyzed by X-ray diffraction combined with microstructural studies and static lattice simulations. The characterization included the measurements of electrical conductivity and oxygen permeability and preliminary assessment of the electrochemical performance of La2NiO4+δ-based electrodes in contact with apatite-type lanthanum silicate-based electrolytes by electrochemical impedance spectroscopy.

2. Materials and Methods

The powders with the nominal composition La2Ni1−ySiyO4+δ (y = 0, 0.02 and 0.05) were synthesized by the glycine-nitrate combustion technique using La(NO3)3·6H2O (99.9%, Sigma-Aldrich, St. Louis, Mo, USA), Ni (99.7%, Sigma-Aldrich), SiO2 (extra pure, Merck) and glycine (≥99%, Sigma Aldrich) as the starting chemical. Ni was dissolved in a minimum amount of nitric acid to yield nickel nitrate. The calculated amount of highly dispersed silica was added to the aqueous solution containing appropriate amounts of lanthanum and nickel nitrates and glycine (glycine/nitrate molar ratio was double with respect to stoichiometric reaction). After stirring for several hours, the homogeneous suspension was heated on a hot plate until evaporation of water and auto-ignition. The foam-like combustion products were fired 3 times at 950 °C/5 h with intermediate regrinding to burn out organic residues and attain phase homogeneity. The as-synthesized powders were ball milled with ethanol (Retsch S1 mill, nylon vial, tetragonal zirconia balls) for 4 h at 150 rpm. After drying, the samples with x = 0.05 were subjected to heat treatments in the powdered or pelletized form to evaluate possible changes in the phase composition with temperature. Final ceramic samples of all materials were compacted uniaxially in a disk-shaped form and sintered at 1450 °C for 5 h.
The density of prepared ceramics (ρexp) was calculated from the geometric dimensions and mass of polished samples. Rectangular bars (~1.5 × 2.5 × 12 mm) for electrical measurements were cut out of disk-shaped samples. Powdered samples for X-ray diffraction (XRD) were prepared by grinding sintered ceramics in a mortar. XRD data were recorded on a Rigaku D/Max-B diffractometer (CuKα radiation). The unit cell parameters were calculated in FullProf software (profile matching method). Microstructural characterization was conducted by scanning electron microscopy (SEM) using a Hitachi SU-70 microscope equipped with Bruker Quantax 400 detector for energy dispersive spectroscopy (EDS) analysis.
Electrical conductivity (σ) was measured as a function of temperature and oxygen partial pressure p(O2) by the four-probe DC technique using bar-shaped samples. Pt wires were used as probes and current collectors. The end-face surfaces of the bars were covered with Pt paint (Heraeus CL11-5349) to improve the electrical contact. The oxygen partial pressure was imposed by the composition of flowing N2+O2 gas mixtures set using Bronkhorst mass-flow controllers. The values of p(O2) were monitored employing a homemade potentiometric yttria-stabilized zirconia (YSZ) sensor. The measurements of oxygen permeation fluxes through dense ceramic membranes were performed at 750–950 °C using electrochemical YSZ solid electrolyte cells comprising an oxygen pump and a sensor [41,42]. The oxygen partial pressure at the membrane feed side (p2) was equal to 0.21 atm (atmospheric air). The thickness of the disk-shaped membranes (d) was 1.00 mm. The gas tightness of the samples before the measurements was verified by the absence of physical leakages under the total pressure gradient of 2–3 atm at room temperature.
The electrochemical characterization of La2NiO4+δ-based electrodes in contact with a La3.83Pr6Si4.5Fe1.5O26 solid electrolyte was performed using electrochemical cells with a symmetrical electrode|electrolyte|electrode configuration. The electrolyte material was synthesized as described in [43]. Dense disk-shaped solid electrolyte membranes with a theoretical density of ~96% were sintered at 1650 °C for 15 h. The electrode ink was prepared by adding as-synthesized La2NiO4+δ powder (30 vol.%) to a solution containing a binder (ethyl cellulose, 5 wt.%) and a dispersant (stearic acid, 3 wt.%) in α-terpineol and ball-milling for several hours. Porous La2NiO4+δ electrode layers (diameter 5 mm) were painted symmetrically on both sides of polished electrolyte pellets and sintered at 1300 °C for 2 h. The electrochemical activity of electrodes was evaluated by electrochemical impedance spectroscopy (EIS) using an AUTOLAB PGSTAT 302 potentiostat/galvanostat with a built-in FRA module (AC amplitude 50 mV, frequency range 10 mHz–1 MHz). The surface modification of electrode layers was performed by infiltration of an aqueous solution of lanthanum nitrate into the porous electrode structure followed by calcination at 1000 °C for 2 h.
The simulations of the La2NiO4+δ lattice were performed in atomistic software using phenomenological potentials. GULP code [44,45] was used for geometry optimization and the evaluation of the lattice energy. In order to obtain the lattice energy of apatite-type silicate lattice, a preliminary generated supercell was treated employing Monte-Carlo simulations in DL_MONTE2 software [46,47] to reach reasonable distribution of ions between partially populated sites in the lanthanum and oxygen sublattices. The most stable configuration obtained in the Monte-Carlo run was selected for the structure optimization in GULP. GULP and DL_MONTE share compatible sets of interatomic potentials that describe interactions of ions with formal charges in the crystalline lattice including the electrostatic term:
U i j C o u l o m b = q i q j 4 π ε 0 r i j ,
where rij is the distance between ions with the charges qi and qj, and the Buckingham term:
U i j B u c k h = A exp ( r i j ρ ) C 6 r i j 6 ,
where A, ρ and C6 are the constants characteristic for each pair of ions. An additional three-body potential was employed for the O2−-Si4+-O2− bonds:
U i j 3 b = 1 2 k 3 b ( θ θ 0 ) 2 ,
where k3b is the angle spring constant, θ0 is the characteristic angle for a given triplet, and θ is the actual angle. The core–shell model was used to account for the polarizability of La3+ and O2− cations:
E i p o l = 1 2 k s p r 2 ,
where r is the shift of the shell relative to the core and ksp is the spring constant. The parameters of phenomenological potentials were collected from the published data and are listed in Table 1. The potential cut-off during the simulations was 18 Å for GULP and 12 Å for DL_MONTE2.

3. Results

XRD analysis of the y = 0.05 powder after firing at 1100 °C (three times for 5 h with intermediate regrinding) revealed the presence of phase impurities in addition to the main tetragonal K2NiF4-type phase (Figure 1A). Taking into account the results of SEM/EDS analysis (discussed below), most extra reflections were assigned to an apatite-type La10−xSi6O26+δ silicate phase. This phase could not be eliminated by increasing the temperature of thermal treatment to 1350–1450 °C and its fraction remained virtually unchanged (Figure 1B–D). Tiny additional impurity peaks on the background level coincide with the positions of the main reflections of La2O3. (Figure 1A). The presence of this phase impurity was detectable only in selected XRD patterns.
The final ceramic samples of all compositions were sintered at 1450 °C for 5 h. The parent silicon-free nickelate was phase pure. The other two compositions contained La10−xSi6O26+δ impurity and, apparently, trace amounts of La2O3 (Figure 2). The fraction of lanthanum silicate phase increased with increasing the nominal silicon content.
The XRD patterns of all compositions were indexed in the tetragonal I4/mmm space group. Additions of silicon have only a minor effect on the lattice parameters of the tetragonal K2NiF4-type lattice (Table 2). Still, the addition of silica results in a slight shrinkage in the basal a-b plane and a modest elongation along the c-axis. This could be explained by the incorporation of higher-valence small Si4+ cations (rVI(Si) = 0.40 Å [51]) into the Ni sublattice (rVI(Ni2+) = 0.69 Å [51]), with expansion along the c-axis caused by incorporation of extra interstitial oxygen required to maintain the electroneutrality condition.
Similar to the parent La2NiO4+δ ceramics, the y = 0.02 samples were gas-tight with comparatively low porosity (Figure 3A) and a relative density of 95% of the theoretical (Table 2). Increasing silica additions resulted in noticeably worse sinterability under identical conditions. The ceramics with the nominal composition La2Ni0.95Si0.05O4+δ had a substantial porosity (Figure 3C) and a relative density of only ~80%. Apart from the pores, the microstructural studies revealed the presence of well-distributed inclusions (appearing with a darker contrast in the SEM images) in both silicon-containing materials (Figure 3). The size of inclusions and their concentration were found to increase with increasing silica additions.
Despite the visual similarities, SEM/EDS inspections showed that there are two types of inclusions. One example is shown in Figure 4. The distribution of lanthanum seemed to be comparatively homogeneous. The inclusions of the first type are enriched with silicon and depleted with nickel. These are the most frequent inclusions corresponding to the La2−xSi6O26+δ impurity phase. More rare inclusions (spot 2 in Figure 4) are depleted in both nickel and silicon and were assigned to La2O3 impurity.
Thus, the results of XRD and SEM/EDS studies imply that the solubility of silicon in the nickel sublattice of La2NiO4+δ is extremely limited under applied synthesis conditions. It is much less than 2 at.% and tends to zero. Instead, silica reacts with lanthanum oxide during the synthesis to yield an apatite-type lanthanum silicate phase. Trace amounts of lanthanum oxide segregate, apparently, to compensate for La:Si < 2 ratio in the apatite-type phase.
As the phase composition of oxide materials can be strongly influenced by the synthetic route, the possibility of the formation of a La2NiO4-based solid solution containing silicon ions in the nickel sublattice was assessed by the static lattice simulations. La 2 3 + Ni 2 + O 4 2 was considered as a host lattice, with all nickel in a 2+ oxidation state. The simulated supercells, their dimensions and lattice energy are listed in Table 3. The latter is a value calculated in GULP as a sum of all energy terms for interatomic interactions for a given oxide. The simulated supercell of silicon-substituted lanthanum nickelate had the full formula La72Ni35SiO145 which can be reduced to approximately La2Ni0.972Si0.028O4.028.
It is necessary to note that the replacement of a nickel cation by a silicon ion should be accompanied by the incorporation of three oxygen ions into the interstitial positions in the rock-salt-type layers of the Ruddlesden-Popper La2NiO4 structure. One oxygen ion is required to keep the lattice’s electroneutrality and compensate for the excessive oxidation state of Si4+ with respect to the host Ni2+ matrix. Two other oxygen anions must be pushed out of the perovskite layer to maintain a tetrahedral oxygen environment strongly preferred by silicon. Thus, the existence of three interstitial oxygen anions in the La2Ni0.972Si0.028O4.028 supercell was assumed. Three configurations involving Si-centered tetrahedron and interstitial O2− ions were tested to address the impact of interactions between charged point defects: (i) two O i 2 placed near Si and one on a sufficient distance; (ii) only one O i 2 placed near Si; and (iii) all O i 2 anions placed on a sufficient distance from the silicon cation. Contrary to the second and third configurations, the first configuration was found to be unstable (as was indicated by the presence of negative lines in the simulated phonon spectrum). The configuration (iii) of the La2Ni0.972Si0.028O4.028 supercell after the geometry optimization is presented in Figure 5.
The energy of the formation of hypothetical La2Ni0.972Si0.028O4.028 was calculated as:
Δ E f = E ( La 2 Ni 0.972 Si 0.028 O 4.028 ) 1 162 E ( La 2 O 3 ) 1 216 E ( La 9.33 Si 6 O 26 ) 35 36 E ( La 2 NiO 4 ) ,
with the coefficients balanced to maintain the cation ratio. Independently of the tested Si - O i 2 configuration, the calculations yielded nearly identical positive values of the formation energy for silicon-substituted lanthanum nickelate (Table 3). This implies that the segregation of apatite-type silicate and lanthanum oxide is energetically more favorable compared to the formation of a solid solution, in agreement with the experimental results.
Inspection of polished and thermally etched ceramic samples by SEM revealed that additions of silica strongly suppress the grain growth during the sintering. In particular, it was found that grain size varied between 3.5 and 20 µm in undoped La2NiO4+δ and decreased down to 0.8–4.0 µm for the y = 0.02 samples prepared under identical conditions. This seems to imply that the silicon-rich phase impurities not only precipitate as individual grains, but also may segregate at the grain boundaries inhibiting the mobility of grain boundaries during the sintering process due to the solute drag effect or Zener effect [52]. This also explained the worse sinterability (higher porosity) of the y = 0.05 ceramics.
Figure 6 presents the data on the electrical conductivity of La2Ni1−ySiyO4+δ ceramics. All compositions exhibit a similar shape of σT curves in air (Figure 6A). Electrical conductivity is temperature activated in the low-temperature range and shows a smooth transition to metallic-like behavior at temperatures above ~400–600 °C, in agreement with literature data on parent La2NiO4+δ [30,53,54]. Additions of silica result in a gradual decrease in electrical conductivity: from 87 S/cm for y = 0 to 68 S/cm for y = 0.02 and to 36 S/cm for y = 0.05 at 800 °C. Note that La10−xSi6O26+δ-based solid electrolytes demonstrate total electrical conductivity several orders of magnitude lower compared to La2NiO4+δ. For instance, the total conductivity of La10Si5.5Al0.5O26+δ, one of the apatite-type silicates with the highest conductivity, corresponds to 5 × 10−2 S/cm at 800 °C [10]. Thus, additions of silica seem to play a rather passive role in terms of the mechanism of electrical transport: electrical conductivity decreases with an increasing fraction of poorly conducting inclusions (and possibly grain boundaries) and pores, as generally expected for different kinds of materials [55]. One should also note that the values of electrical conductivity of La2Ni0.98Si0.02O4+δ obtained in this work are comparable or even exceed the values often reported in the literature for undoped La2NiO4+δ (e.g., refs. [53,54]); most likely, this should be assigned to the differences in porosity and microstructure.
Reducing oxygen partial pressure results in a decrease in the electrical conductivity of La2Ni1−ySiyO4+δ ceramics (Figure 6B), in agreement with the literature data for undoped lanthanum nickelate [56,57]. This is associated with the reversible oxygen release from the K2NiF4-type lattice accompanied by the reduction in nickel cations and, therefore, a decrease in the concentration of electronic charge carriers:
2 Ni Ni + O i p O 2 p O 2 2 Ni Ni × + 0.5 O 2 ,
where Ni Ni and Ni Ni × correspond to Ni3+ and Ni2+, respectively, and Ni Ni is equivalent to electron-hole.
Figure 7 shows the data on oxygen permeability of La2NiO4+δ and La2Ni0.98Si0.02O4+δ ceramic membranes. These measurements could not be performed for the material with higher silica content since La2Ni0.95Si0.05O4+δ ceramics had excessive porosity and were not gas tight. Additions of silica to La2NiO4+δ were found to suppress oxygen ionic transport through ceramic membranes. Similar to the case of electrical conductivity, this can be partially attributed to the increasing fraction of the poorly conducting phase in the bulk of ceramics impeding the oxygen transport. However, the drop in oxygen permeability with additions of silica is accompanied by an increase in activation energy of the permeation process. This can possibly be caused by one of two following factors. Assuming partial incorporation of silicon into Ni sites, one may expect strong trapping of mobile ionic species (interstitial oxygen ions) near the Si4+ cations with effects on overall oxygen-ion mobility and activation energy of ionic transport. For instance, the substitution of iron with silicon was reported to suppress oxygen-ionic conductivity in perovskite-type SrFe1−ySiyO4−δ solid solutions [36]; this was explained by anion site exclusion effects and oxygen vacancy trapping near SiO4 tetrahedra [58]. However, since the solubility of silicon in La2NiO4+δ lattices is extremely limited (if any), the more probable reason for the drop in oxygen permeability is inhibition of surface exchange kinetics due to partial blocking of the surface by silica species. Increasing surface exchange limitations by silica impurities or silica poisoning were reported earlier for other mixed conducting ceramic membranes (e.g., refs. [59,60]).
Thus, the presence of silica is unfavorable for the surface processes involving oxygen and, therefore, may inhibit the electrochemical activity of mixed-conducting ceramic electrodes. This is in agreement with literature data suggesting that, even though Ln2NiO4+δ electrodes show good chemical compatibility with apatite-type silicate electrolytes [19,23,26], the deterioration of electrochemical activity of electrodes occurs due to the surface diffusion or spreading of silica from electrolyte to electrode resulting in a partial blocking of the electrochemical reaction zone [22,23,24,26]. However, since silicon is practically insoluble in the La2NiO4+δ lattice and the preferential reactivity pathway is the formation of the La10−xSi6O26+δ-based phase, this opens the opportunity to minimize the negative effects of silica diffusion from the silicate solid electrolyte onto the surface of La2NiO4-based electrodes by additions of extra lanthanum oxide to the electrode. As a result, the reactivity between lanthanum oxide and diffusing silica may result in the formation of solid electrolyte particles thus extending the triple-phase boundary instead of blocking the electrode active surface. A similar approach was proposed earlier for silica scavenging in contaminated zirconia and ceria solid electrolytes [61,62].
The preliminary tests seem to support the potential efficiency of such an approach. Figure 8 compares the electrochemical impedance spectra of La2NiO4+δ applied onto apatite-type La3.83Pr6Si4.5Fe1.5O26 solid electrolyte before and after infiltration. The spectra can be roughly fitted using a simple equivalent circuit (ROhm)(RHF||CPEHF)(RLF||CPELF) where R is resistance, CPE is the constant phase element, ROhm is the ohmic resistance of the cell, HF and LF indicate high-frequency and low-frequency contribution to the electrode process, and total electrode polarization resistance is the sum of RHF+RLF. Two electrode contributions were estimated to have relaxation frequencies of around 8–9 kHz and 100–300 Hz, respectively. Although these preliminary impedance spectroscopy data are not sufficient for reliable conclusions about the impact of surface modification on the different stages of the electrode process, it is evident that the infiltration of lanthanum oxide resulted in a drop in electrode polarization resistance by ~1.5 times at 800 °C. The impact was lower when the temperature was increased to 900 °C, possibly due to enhanced silica diffusion at a higher temperature. Nonetheless, these results support the potential suitability of the method. Further detailed studies are planned to explore the impact of surface modifications (thermal treatment procedure, infiltrated fraction) on the electrode performance, limiting stages of electrode process, and long-term stability.

4. Conclusions

Ceramics with the nominal composition La2Ni1−ySiyO4+δ (y = 0, 0.02 and 0.05) were prepared by the glycine-nitrate combustion method and sintered at 1450 °C. Minor changes in the a and c parameters of tetragonal K2NiF4-type lattice with increasing y suggest possible partial incorporation of silicon cations into the nickel sublattice. However, the results of combined XRD and SEM/EDS studies indicate that the solubility of silicon in the La2NiO4+δ lattice is very limited (≪2 at.% in Ni sublattice, if any). Instead, additions of silica to La2NiO4+δ induce the precipitation of apatite-type La10−xSi6O26+δ phase and traces of La2O3. The preferential formation of secondary phases in place of La2Ni1−ySiyO4+δ solid solutions is supported by the static lattice simulations. The segregation of phase impurities suppresses the grain growth and results in a gradual decline in electrical conductivity and oxygen permeability of La2NiO4+δ ceramics with the increasing silicon content. Practical insolubility of silicon in the lanthanum nickelate lattice and preferential formation of apatite-type lanthanum silicate phase suggest the possibility to inhibit the degradation of La2NiO4-based electrodes in contact with La10−xSi6O26+δ solid electrolytes by additions of extra La2O3 to capture diffusing silica. The prospects of such an approach are supported by preliminary tests and will be explored in further work.

Author Contributions

Conceptualization and methodology, A.A.Y.; static lattice simulations, E.N.N.; investigation, K.Z., A.B., M.A.V. and A.A.Y.; writing—original draft preparation, A.A.Y. and E.N.N.; writing—review and editing, K.Z., A.B., E.N.N., M.A.V. and A.A.Y.; funding acquisition, A.A.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the project CARBOSTEAM (POCI-01-0145-FEDER-032295) funded by FEDER through COMPETE2020—Programa Operacion-al Competitividade e Internacionalização (POCI) and by national funds through FCT/MCTES, and by the project CICECO—Aveiro Institute of Materials (UIDB/50011/2020 and UIDP/50011/2020) financed by national funds through the FCT/MCTES and when appropriate co-financed by FEDER under the PT2020 Partnership Agreement. PhD scholarship of K.Z. is funded by the FCT (grant SFRH/BD/138773/2018). PhD scholarship of A.B. is funded by the FCT (grant SFRH/BD/150704/2020).

Data Availability Statement

Data are contained within the article and/or available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Singh, M.; Zappa, D.; Comini, E. Solid oxide fuel cell: Decade of progress, future perspectives and challenges. Int. J. Hydrogen Energy 2021, 46, 27643–27674. [Google Scholar] [CrossRef]
  2. Cigolotti, V.; Genovese, M.; Fragiacomo, P. Comprehensive review on fuel cell technology for stationary applications as sustainable and efficient poly-generation energy systems. Energies 2021, 14, 4963. [Google Scholar] [CrossRef]
  3. Vostakola, M.F.; Horri, B.A. Progress in material development for low-temperature solid oxide fuel cells: A review. Energies 2021, 14, 1280. [Google Scholar] [CrossRef]
  4. Jacobson, A.J. Materials for solid oxide fuel cells. Chem. Mater. 2010, 22, 660–674. [Google Scholar] [CrossRef]
  5. Mineshige, A.; Saito, A.; Kobayashi, M.; Hayakawa, H.; Momai, M.; Yazawa, T.; Yoshioka, H.; Sakao, M.; Mori, R.; Takayama, Y.; et al. Lanthanum silicate-based layered electrolyte for intermediate-temperature fuel cell application. J. Power Sources 2020, 475, 228543. [Google Scholar] [CrossRef]
  6. Wang, S.F.; Hsu, Y.F.; Hsia, P.; Hung, W.K.; Jasinski, P. Design and characterization of apatite La9.8Si5.7Mg0.3O26±δ-based micro-tubular solid oxide fuel cells. J. Power Sources 2020, 460, 228072. [Google Scholar] [CrossRef]
  7. Marrero-López, D.; Martín-Sedeño, M.C.; Peña-Martínez, J.; Ruiz-Morales, J.C.; Núñez, P.; Aranda, M.A.G.; Ramos-Barrado, J.R. Evaluation of apatite silicates as solid oxide fuel cell electrolytes. J. Power Sources 2010, 195, 2496–2506. [Google Scholar] [CrossRef]
  8. Malavasi, L.; Fisher, C.A.J.; Islam, M.S. Oxide-ion and proton conducting electrolyte materials for clean energy applications: Structural and mechanistic features. Chem. Soc. Rev. 2010, 39, 4370–4387. [Google Scholar] [CrossRef] [PubMed]
  9. Kendrick, E.; Islam, M.S.; Slater, P.R. Developing apatites for solid oxide fuel cells: Insight into structural, transport and doping properties. J. Mater. Chem. 2007, 17, 3104–3111. [Google Scholar] [CrossRef]
  10. Shaula, A.L.; Kharton, V.V.; Marques, F.M.B. Oxygen ionic and electronic transport in apatite-type La10-x(Si,Al)6O27+δ. J. Solid State Chem. 2005, 178, 2050–2061. [Google Scholar] [CrossRef]
  11. Kharton, V.V.; Shaula, A.L.; Patrakeev, M.V.; Waerenborgh, J.C.; Rojas, D.P.; Vyshatko, N.P.; Tsipis, E.V.; Yaremchenko, A.A.; Marques, F.M.B. Oxygen ionic and electronic transport in apatite-type solid electrolytes. J. Electrochem. Soc. 2004, 151, A1236–A1246. [Google Scholar] [CrossRef]
  12. Yoshioka, H.; Nojiri, Y.; Tanase, S. Ionic conductivity and fuel cell properties of apatite-type lanthanum silicates doped with Mg and containing excess oxide ions. Solid State Ion. 2008, 179, 2165–2169. [Google Scholar] [CrossRef]
  13. Mineshige, A.; Hayakawa, H.; Nishimoto, T.; Heguri, A.; Yazawa, T.; Takayama, Y.; Kagoshima, Y.; Takano, H.; Takeda, S.; Matsui, J. Preparation of lanthanum silicate electrolyte with high conductivity and high chemical stability. Solid State Ion. 2018, 319, 223–227. [Google Scholar] [CrossRef]
  14. Nakayama, S.; Aono, H.; Sadaoka, Y. Ionic conductivity of Ln10(SiO4)6O3 (Ln = La, Nd, Sm, Gd and Dy). Chem. Lett. 1995, 431–432. [Google Scholar] [CrossRef]
  15. Yaremchenko, A.A.; Kovalevsky, A.V.; Kharton, V.V. Mixed conductivity, stability and electrochemical behavior of perovskite-type (Sr0.7Ce0.3)1-xMn1-yCryO3-δ. Solid State Ion. 2008, 179, 2181–2191. [Google Scholar] [CrossRef]
  16. Yaremchenko, A.A.; Bannikov, D.O.; Kovalevsky, A.V.; Cherepanov, V.A.; Kharton, V.V. High-temperature transport properties, thermal expansion and cathodic performance of Ni-substituted LaSr2Mn2O7-δ. J. Solid State Chem. 2008, 181, 3024–3032. [Google Scholar] [CrossRef]
  17. Brisse, A.; Sauvet, A.L.; Barthet, C.; Beaudet-Savignat, S.; Fouletier, J. Electrochemical characterizations of Ni/doped lanthanum silicates cermets deposited by spin coating. Fuel Cells 2006, 6, 59–63. [Google Scholar] [CrossRef]
  18. Gasparyan, H.; Argirusis, C.; Szepanski, C.; Sourkouni, G.; Stathopoulos, V.; Kharlamova, T.; Sadykov, V.; Bebelis, S. Electrochemical characterization of a La0.8Sr0.2Ni0.4Fe0.6O3-δ electrode interfaced with La9.83Si5Al0.75Fe0.25O26±δ apatite-type electrolyte. ECS Trans. 2009, 25, 2681–2688. [Google Scholar] [CrossRef]
  19. Philippeau, B.; Mauvy, F.; Mazataud, C.; Fourcade, S.; Grenier, J.C. Comparative study of electrochemical properties of mixed conducting Ln2NiO4+δ (Ln = La, Pr and Nd) and La0.6Sr0.4Fe0.8Co0.2O3-δ as SOFC cathodes associated to Ce0.9Gd0.1O2-δ, La0.8Sr0.2Ga0.8Mg0.2O3-δ and La9Sr1Si6O26.5 electrolytes. Solid State Ion. 2013, 249-250, 17–25. [Google Scholar] [CrossRef]
  20. Zhou, J.; Ye, X.F.; Li, J.L.; Wang, S.R.; Wen, T.L. Synthesis and characterization of apatite-type La9.67Si6-xAlxO26.5-x/2 electrolyte materials and compatible cathode materials. Solid State Ion. 2011, 201, 81–86. [Google Scholar] [CrossRef]
  21. Bonhomme, C.; Beaudet-Savignat, S.; Chartier, T.; Geffroy, P.M.; Sauvet, A.L. Evaluation of the La0.75Sr0.25Mn0.8Co0.2O3-δ system as cathode material or IT-SOFCs with La9Sr1Si6O26.5 apatite as electrolyte. J. Eur. Ceram. Soc. 2009, 29, 1781–1788. [Google Scholar] [CrossRef]
  22. Tsipis, E.V.; Kharton, V.V.; Frade, J.R. Electrochemical behavior of mixed-conducting oxide cathodes in contact with apatite-type La10Si5AlO26.5 electrolyte. Electrochim. Acta 2007, 52, 4428–4435. [Google Scholar] [CrossRef]
  23. Yaremchenko, A.A.; Kharton, V.V.; Bannikov, D.O.; Znosak, D.V.; Frade, J.R.; Cherepanov, V.A. Performance of perovskite-related oxide cathodes in contact with lanthanum silicate electrolyte. Solid State Ion. 2009, 180, 878–885. [Google Scholar] [CrossRef]
  24. Cao, X.G.; Jiang, S.P. Identification of oxygen reduction processes at (La,Sr)MnO3 electrode/La9.5Si6O26.25 apatite electrolyte interface of solid oxide fuel cells. Int. J. Hydrogen Energy 2013, 38, 2421–2431. [Google Scholar] [CrossRef]
  25. McFarlane, J.; Barth, S.; Swaffer, M.; Sansom, J.E.H.; Slater, P.R. Synthesis and conductivities of the apatite-type systems, La9.33+xSi6-yMyO26+z (M = Co, Fe, Mn) and La8Mn2Si6O26. Ionics 2002, 8, 149–154. [Google Scholar] [CrossRef]
  26. Antonova, E.P.; Osinkin, D.A.; Bogdanovich, N.M.; Gorshkov, M.Y.; Bronin, D.I. Electrochemical performance of Ln2NiO4+δ (Ln–La, Nd, Pr) and Sr2Fe1.5Mo0.5O6-δ oxide electrodes in contact with apatite-type La10(SiO6)4O3 electrolyte. Solid State Ion. 2019, 329, 82–89. [Google Scholar] [CrossRef]
  27. Robles-Fernandez, A.; Orera, A.; Merino, R.I.; Slater, P.R. Suitability of strontium and cobalt-free perovskite cathodes with La9.67Si5AlO26 apatite electrolyte for intermediate temperature solid oxide fuel cells. Dalton Trans. 2020, 49, 14280–14289. [Google Scholar] [CrossRef]
  28. Das, A.; Xhafa, E.; Nikolla, E. Electro- and thermal-catalysis by layered, first series Ruddlesden-Popper oxides. Catal. Today 2016, 277, 214–226. [Google Scholar] [CrossRef] [Green Version]
  29. Xu, X.; Pan, Y.; Zhong, Y.; Ran, R.; Shao, Z. Ruddlesden-Popper perovskites in electrocatalysis. Mater. Horiz. 2020, 7, 2519–2565. [Google Scholar] [CrossRef]
  30. Tarutin, A.P.; Lyagaeva, J.G.; Medvedev, D.A.; Bi, L.; Yaremchenko, A.A. Recent advances in layered Ln2NiO4+δ nickelates: Fundamentals and prospects of their applications in protonic ceramic fuel and electrolysis cells. J. Mater. Chem. A 2021, 9, 154–195. [Google Scholar] [CrossRef]
  31. Lee, D.; Lee, H.N. Controlling oxygen mobility in Ruddlesden-Popper oxides. Materials 2017, 10, 368. [Google Scholar] [CrossRef]
  32. Ding, P.; Li, W.; Zhao, H.; Wu, C.; Zhao, L.; Dong, B.; Wang, S. Review on Ruddlesden-Popper perovskites as cathode for solid oxide fuel cells. J. Phys. Mater. 2021, 4, 022002. [Google Scholar] [CrossRef]
  33. Hancock, C.A.; Porras-Vazquez, J.M.; Keenan, P.J.; Slater, P.R. Oxyanions in perovskites: From superconductors to solid oxide fuel cells. Dalton Trans. 2015, 44, 10559–10569. [Google Scholar] [CrossRef] [Green Version]
  34. Hancock, C.A.; Slater, P.R. Synthesis of silicon doped SrMO3 (M = Mn, Co): Stabilization of the cubic perovskite and enhancement in conductivity. Dalton Trans. 2011, 40, 5599–5603. [Google Scholar] [CrossRef]
  35. Porras-Vazquez, J.M.; Pike, T.; Hancock, C.A.; Marco, J.F.; Berrya, F.J.; Slater, P.R. Investigation into the effect of Si doping on the performance of SrFeO3-δ SOFC electrode materials. J. Mater. Chem. A 2013, 1, 11834–11841. [Google Scholar] [CrossRef] [Green Version]
  36. Merkulov, O.V.; Markova, A.A.; Patrakeev, M.V.; Chukin, A.V.; Leonidov, I.A.; Kozhevnikov, V.L. Structural features and high-temperature properties of SrFe1-xSixO3-δ. Solid State Ion. 2016, 292, 83–87. [Google Scholar] [CrossRef]
  37. Porras-Vazquez, J.M.; Losilla, E.R.; Keenan, P.J.; Hancock, C.A.; Kemp, T.F.; Hanna, J.V.; Slater, P.R. Investigation into the effect of Si doping on the performance of Sr1-yCayMnO3-δ SOFC cathode materials. Dalton Trans. 2013, 42, 5421–5429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Porras-Vazquez, J.M.; Smith, R.I.; Slater, P.R. Investigation into the effect of Si doping on the cell symmetry and performance of Sr1-yCayFeO3-δ SOFC cathode materials. J. Solid State Chem. 2014, 213, 132–137. [Google Scholar] [CrossRef]
  39. Smith, A.D.; James, M.S.; Slater, P.R. Effect of Si-doping on the structure and conductivity of (Sr/Ca)2MnFeO6-δ systems. ECS Trans. 2019, 91, 1425–1436. [Google Scholar] [CrossRef] [Green Version]
  40. Porras-Vazquez, J.M.; Slater, P.R. Synthesis and characterization of oxyanion-doped cobalt containing perovskites. Fuel Cells 2012, 12, 1056–1063. [Google Scholar] [CrossRef]
  41. Kharton, V.V.; Tikhonovich, V.N.; Shuangbao, L.; Naumovich, E.N.; Kovalevsky, A.V.; Viskup, A.P.; Bashmakov, I.A.; Yaremchenko, A.A. Ceramic microstructure and oxygen permeability of SrCo(Fe,M)O3-δ (M = Cu or Cr) perovskite membranes. J. Electrochem. Soc. 1998, 145, 1363–1374. [Google Scholar] [CrossRef]
  42. Kovalevsky, A.V.; Yaremchenko, A.A.; Kolotygin, V.A.; Shaula, A.L.; Kharton, V.V.; Snijkers, F.M.M.; Buekenhoudt, A.; Frade, J.R.; Naumovich, E.N. Processing and oxygen permeation studies of asymmetric multilayer Ba0.5Sr0.5Co0.8Fe0.2O3-δ membranes. J. Membr. Sci. 2011, 380, 68–80. [Google Scholar] [CrossRef]
  43. Yaremchenko, A.A.; Shaula, A.L.; Kharton, V.V.; Waerenborgh, J.C.; Rojas, D.P.; Vyshatko, N.P.; Patrakeev, M.V.; Marques, F.M.B. Ionic and electronic conductivity of La9.83-xPrxSi4.5Fe1.5O26±δ apatites. Solid State Ion. 2004, 171, 51–59. [Google Scholar] [CrossRef]
  44. Gale, J.D. GULP—A computer program for the symmetry adapted simulation of solids. J. Chem. Soc. Faraday Trans. 1997, 93, 629–637. [Google Scholar] [CrossRef]
  45. Gale, J.D.; Rohl, A.L. The general utility lattice program (GULP). Mol. Simul. 2003, 29, 291–341. [Google Scholar] [CrossRef]
  46. Purton, J.A.; Crabtree, J.C.; Parker, S.C. DL_MONTE: A general purpose program for parallel Monte Carlo simulation. Mol. Simul. 2013, 39, 1240–1252. [Google Scholar] [CrossRef]
  47. Brukhno, A.V.; Grant, J.; Underwood, T.L.; Stratford, K.; Parker, S.C.; Purton, J.A.; Wilding, N.B. DL_MONTE: A multipurpose code for Monte Carlo simulation. Mol. Simul. 2021, 47, 131–151. [Google Scholar] [CrossRef] [Green Version]
  48. Tolchard, J.R.; Islam, M.S.; Slater, P.R. Defect chemistry and oxygen ion migration in the apatite-type materials La9.33Si6O26 and La8Sr2Si6O26. J. Mater. Chem. 2003, 13, 1956–1961. [Google Scholar] [CrossRef] [Green Version]
  49. DFRL-UCL Potentials Library: Ni. Available online: https://www.ucl.ac.uk/klmc/Potentials/Ni/index.html (accessed on 16 November 2021).
  50. DFRL-UCL Potentials Library: Silicon. Available online: https://www.ucl.ac.uk/klmc/Potentials/Si/index.html (accessed on 16 November 2021).
  51. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chaleogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  52. Kang, S.J.L. Sintering. Densification, Grain Growth, and Microstructure; Elsevier Butterworth-Heinemann: Amsterdam, The Netherlands, 2005; pp. 95–103. [Google Scholar]
  53. Boehm, E.; Bassat, J.M.; Dordor, P.; Mauvy, F.; Grenier, J.C.; Stevens, P. Oxygen diffusion and transport properties in non-stoichiometric Ln2-xNiO4+δ oxides. Solid State Ion. 2005, 176, 2717–2725. [Google Scholar] [CrossRef]
  54. Pikalova, E.Y.; Kolchugin, A.A.; Sadykov, V.A.; Sadovskaya, E.M.; Filonova, E.A.; Eremeev, N.F.; Bogdanovich, N.M. Structure, transport properties and electrochemical behavior of the layered lanthanide nickelates doped with calcium. Int. J. Hydrogen Energy 2018, 43, 17373–17386. [Google Scholar] [CrossRef]
  55. Ternero, F.; Rosa, L.G.; Urban, P.; Montes, J.M.; Cuevas, F.G. Influence of the total porosity on the properties of sintered materials—A review. Metals 2021, 11, 730. [Google Scholar] [CrossRef]
  56. Shaula, A.L.; Naumovich, E.N.; Viskup, A.P.; Pankov, V.V.; Kovalevsky, A.V.; Kharton, V.V. Oxygen transport in La2NiO4+δ: Assessment of surface limitations and multilayer membrane architectures. Solid State Ion. 2009, 180, 812–816. [Google Scholar] [CrossRef]
  57. Nakamura, T.; Yashiro, K.; Sato, K.; Mizusaki, J. Electronic state of oxygen nonstoichiometric La2-xSrxNiO4+δ at high temperatures. Phys. Chem. Chem. Phys. 2009, 11, 3055–3062. [Google Scholar] [CrossRef]
  58. Merkulov, O.V.; Naumovich, E.N.; Patrakeev, M.V.; Markov, A.A.; Shalaeva, E.V.; Kharton, V.V.; Tsipis, E.V.; Waerenborgh, J.C.; Leonidov, I.A.; Kozhevnikov, V.L. Defect formation, ordering, and transport in SrFe1-xSixO3-δ (x = 0.05–0.20). J. Solid State Electrochem. 2018, 22, 727–737. [Google Scholar] [CrossRef]
  59. Viitanen, M.M.; Welzenis, R.G.v.; Brongersma, H.H.; van Berkel, F.P.F. Silica poisoning of oxygen membranes. Solid State Ion. 2002, 150, 223–228. [Google Scholar] [CrossRef]
  60. Liu, Y.; Zhu, X.; Li, M.; Li, W.; Yang, W. Degradation and stabilization of perovskite membranes containing silicon impurity at low temperature. J. Membr. Sci. 2015, 492, 173–180. [Google Scholar] [CrossRef]
  61. Carvalho, T.; Antunes, E.; Calado, J.; Figueiredo, F.M.; Frade, J.R. Lanthanum oxide as a scavenging agent for zirconia electrolytes. Solid State Ion. 2012, 225, 484–487. [Google Scholar] [CrossRef]
  62. Ivanova, D.; Lima, E.M.C.L.G.P.; Kovalevsky, A.; Figueiredo, F.M.L.; Kharton, V.V.; Marques, F.M.B. Heterogeneous ceramics formed by grain boundary engineering. Ionics 2008, 14, 349–356. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of La2Ni0.95Si0.05O4+δ samples after thermal treatments: (A) powder and (BD) compacted samples sintered at different temperatures.
Figure 1. XRD patterns of La2Ni0.95Si0.05O4+δ samples after thermal treatments: (A) powder and (BD) compacted samples sintered at different temperatures.
Processes 10 00082 g001
Figure 2. XRD patterns of La2Ni1−ySiyO4+δ ceramics sintered at 1450 °C for 5 h.
Figure 2. XRD patterns of La2Ni1−ySiyO4+δ ceramics sintered at 1450 °C for 5 h.
Processes 10 00082 g002
Figure 3. SEM images of polished La2Ni1−ySiyO4+δ ceramics.
Figure 3. SEM images of polished La2Ni1−ySiyO4+δ ceramics.
Processes 10 00082 g003
Figure 4. SEM image of the polished surface of La2Ni0.98Si0.02O4+δ ceramics and corresponding EDS elemental maps. Spots 1 and 2 correspond to La2−xSi6O26+δ and La2O3 impurity phase inclusions, respectively.
Figure 4. SEM image of the polished surface of La2Ni0.98Si0.02O4+δ ceramics and corresponding EDS elemental maps. Spots 1 and 2 correspond to La2−xSi6O26+δ and La2O3 impurity phase inclusions, respectively.
Processes 10 00082 g004
Figure 5. Simulated La2Ni0.972Si0.028O4.028 supercell with configuration (iii) after geometry optimization.
Figure 5. Simulated La2Ni0.972Si0.028O4.028 supercell with configuration (iii) after geometry optimization.
Processes 10 00082 g005
Figure 6. Electrical conductivity of La2Ni1−ySiyO4+δ ceramics as a function of temperature in air (A) and as a function of oxygen partial pressure at 900 °C (B).
Figure 6. Electrical conductivity of La2Ni1−ySiyO4+δ ceramics as a function of temperature in air (A) and as a function of oxygen partial pressure at 900 °C (B).
Processes 10 00082 g006
Figure 7. (A) Dependence of oxygen permeation fluxes through La2Ni0.98Si0.02O4+δ ceramic membrane on oxygen partial pressure gradient at 750–950 °C; (B) comparison of oxygen permeation fluxes through La2Ni1−ySiyO4+δ membranes under fixed p(O2) gradient.
Figure 7. (A) Dependence of oxygen permeation fluxes through La2Ni0.98Si0.02O4+δ ceramic membrane on oxygen partial pressure gradient at 750–950 °C; (B) comparison of oxygen permeation fluxes through La2Ni1−ySiyO4+δ membranes under fixed p(O2) gradient.
Processes 10 00082 g007
Figure 8. Electrochemical impedance spectra of La2NiO4+δ electrode, as prepared and infiltrated with lanthanum oxide, in contact with apatite-type La3.83Pr6Si4.5Fe1.5O26 solid electrolyte at 800 °C (A) and 900 °C (B). The spectra are corrected for ohmic contribution.
Figure 8. Electrochemical impedance spectra of La2NiO4+δ electrode, as prepared and infiltrated with lanthanum oxide, in contact with apatite-type La3.83Pr6Si4.5Fe1.5O26 solid electrolyte at 800 °C (A) and 900 °C (B). The spectra are corrected for ohmic contribution.
Processes 10 00082 g008
Table 1. Interatomic potentials used for the lattice simulation.
Table 1. Interatomic potentials used for the lattice simulation.
InteractionBuckingham TermCore-Shell TermThree-Body TermRef.
A, eVρ, ÅC, Å−6Y, eksp, eV Å−2k3b, eV rad−2θ0, °
O2−-O2−22,764.30.14927.89−2.86074.92 [48]
La3+-O2−4579.230.30437-−0.250145.00 [48]
Ni2+-O2−1582.500.28820--- [49]
Si4+-O2−1283.910.3205210.66-- [48]
O2−-Si4+-O2− 2.097240109.47[50]
Table 2. Lattice parameters and density of La2Ni1−ySiyO4+δ ceramics sintered at 1450 °C.
Table 2. Lattice parameters and density of La2Ni1−ySiyO4+δ ceramics sintered at 1450 °C.
CompositionUnit Cell ParametersDensity
ρexp, g/cm3
Relative Density
ρexptheor, % 1
a, Åc, ÅV, Å3
03.8636(1)12.6662(2)189.07(1)6.8296.3
0.023.8611(1)12.6784(4)189.02(1)6.7395.2
0.053.8610(1)12.6831(4)189.08(1)5.6079.5
1 ρtheor was estimated from the structural data assuming nominal composition with δ = 0.15 and neglecting phase impurities.
Table 3. Simulated supercells and calculated lattice energies per formula unit.
Table 3. Simulated supercells and calculated lattice energies per formula unit.
Cation CompositionSupercellLattice Energy, eV/f.u.Formation Energy
Δ E f ,   eV / f . u .
La2O31 × 1 × 1−129.92
La2NiO43 × 3 × 2−171.40
La9.33Si6O263 × 3 × 3−1395.06
La2Ni0.972Si0.028O4.0283 × 3 × 2, config.(i)−173.66 10.24
3 × 3 × 2, config.(ii)−173.640.26
3 × 3 × 2, config.(iii)−173.610.29
1 Unstable configuration.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zakharchuk, K.; Bamburov, A.; Naumovich, E.N.; Vieira, M.A.; Yaremchenko, A.A. Impact of Silica Additions on the Phase Composition and Electrical Transport Properties of Ruddlesden-Popper La2NiO4+δ Mixed Conducting Ceramics. Processes 2022, 10, 82. https://doi.org/10.3390/pr10010082

AMA Style

Zakharchuk K, Bamburov A, Naumovich EN, Vieira MA, Yaremchenko AA. Impact of Silica Additions on the Phase Composition and Electrical Transport Properties of Ruddlesden-Popper La2NiO4+δ Mixed Conducting Ceramics. Processes. 2022; 10(1):82. https://doi.org/10.3390/pr10010082

Chicago/Turabian Style

Zakharchuk, Kiryl, Aleksandr Bamburov, Eugene N. Naumovich, Miguel A. Vieira, and Aleksey A. Yaremchenko. 2022. "Impact of Silica Additions on the Phase Composition and Electrical Transport Properties of Ruddlesden-Popper La2NiO4+δ Mixed Conducting Ceramics" Processes 10, no. 1: 82. https://doi.org/10.3390/pr10010082

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

Article Metrics

Back to TopTop