Next Article in Journal
Comparative Analysis of Strength and Deformation Behavior of Cemented Tailings Backfill under Curing Temperature Effect
Next Article in Special Issue
Molten Chlorides as the Precursors to Modify the Ionic Composition and Properties of LiNbO3 Single Crystal and Fine Powders
Previous Article in Journal
Prediction of Mechanical Properties of Fly-Ash/Slag-Based Geopolymer Concrete Using Ensemble and Non-Ensemble Machine-Learning Techniques
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Layered Perovskites BaM2In2O7 (M = La, Nd): From the Structure to the Ionic (O2−, H+) Conductivity

Institute of High Temperature Electrochemistry of the Ural Branch of the Russian Academy of Sciences, 620066 Ekatherinburg, Russia
*
Author to whom correspondence should be addressed.
Materials 2022, 15(10), 3488; https://doi.org/10.3390/ma15103488
Submission received: 15 April 2022 / Revised: 9 May 2022 / Accepted: 11 May 2022 / Published: 12 May 2022
(This article belongs to the Special Issue Electrochemical Processes, Materials and Devices)

Abstract

:
The design of new oxide compounds that can be used as oxygen- or proton-conducting electrolytes for solid oxide fuel cells is actively in progress. Despite the intensive research activities regarding electrolytes with perovskite/fluorite structures, the search for other structural alternatives is of paramount importance. In this study we focus on a novel material with significantly improved properties for the electrochemical purposes. The two-layered BaNd2In2O7 perovskite with a Ruddlesden–Popper structure was investigated as a protonic conductor for the first time. In detail, its local structure, water uptake, and the ionic (O2−, H+) conductivity were comprehensively studied. The nature of rare-earth elements (M = La, Nd) in the structure of BaM2In2O7 on the structural and transport properties was revealed. The presented analysis showed that the composition of BaNd2In2O7 is nearly pure proton conductor below 350 °C. This work opens up a new way in the design of protonic conductors with double-layered perovskite structure.

1. Introduction

The materials with layered perovskite-related structures have many various applications due to their different physical-chemical properties. These properties are dependent on the nature of ions in the crystal lattice. The structure of the compositions with the general formula AA 2 B 2 O 7 or AO(A′BO3)2 can be described as the block-layered Ruddlesden–Popper (RP) structure where blocks consisting of two perovskite octahedra A′BO3 alternate with the salt AO layers [1]. The sum of charges of the cations in the A/A′- and B-sublattices can be obtain by the different variations such as +4 and +10, +6 and +8, and +8 and +6 (Figure 1). The total charge +4 in the A/A′-sublattice is typical for photocatalysts A2A′Ta2O7 where A is the hydrogen or alkali metal and A′ is the alkali-earth metal [2,3,4]. The phosphors such as Sr3Sn2O7:Eu3+ [5] and Sr3Ti2O7:Eu3+ [6] have total charge +6 and +8 in the A- and B-sublattices, respectively. For the magnetic AIILn2Mn2O7 [7,8,9,10,11,12] and cathode AIILn2MM′O7 [13,14,15,16] materials the sum of cationic charges can be written as +8 and +6, where AII is the alkali-earth metal, Ln is the lanthanide, M and M′ are metals with variable oxidation state such as Mn, Fe, Co, or Ni. In the case of the presence in the B-sublattice the metal with constant oxidation state such as indium, the chemical properties become completely different despite the same +8/+6 sum of cationic charges. As it was shown recently [17], the composition BaLa2In2O7 demonstrates nearly pure protonic transport under wet air and low temperatures and this phase can be potentially considered as the electrolytic material for the solid oxide fuel cell; therefore, the development of new materials with improvement properties is very relevant today [18,19,20,21,22,23,24,25,26].
The structure of the compounds BaM2In2O7 (M = La, Nd) was described earlier by Titov et al. [27] and Raveau et al. [28]. It was shown that these phases belong to the RP structure AO(A′BO3)n where n = 2. The compositions with monolayered (n = 1) RP structure in the system Ba-Ln-In-O are exist also. The complex oxides BaNdInO4 [29,30,31,32,33,34] and BaLaInO4 [35,36,37,38,39,40,41] and doped compositions based on it are mixed ionic-electronic or protonic conductors depending on the water partial pressure in the atmosphere and temperature. At the same time, in the row BaLaInO4—BaNdInO4 the ionic conductivity increased. Based on this, it can be predicted the same dependency in the row BaLa2In2O7–BaNd2In2O7. From this point of view, the phase BaNd2In2O7 is of interest for studying their physicochemical properties.
In this paper, the local structure, water uptake, and the ionic (O2−, H+) transport in the complex oxide BaNd2In2O7 were investigated for the first time. It was shown that this material can be considered as the promising matrix compositions for development of novel high-conductive protonic electrolytes with PR structure.

2. Materials and Methods

The complex oxides BaNd2In2O7 and BaLa2In2O7 were obtained by the method of solid-state synthesis. The powder initial materials BaCO3, La2O3, Nd2O3, and In2O3 were previously dried, weighed, and mixed in stoichiometric quantities. The milling of powders was made in agate mortar. The compositions were calcined at 800, 900, 1000, 1100, 1200, and 1300 °C for 24 h at air, intermediate grindings were made for every following heating step.
The X-ray diffraction studies were made by the Bruker Advance D8 Cu Kα diffractometer (Billerica, MA, USA) with a step of 0.01° and at a scanning rate of 0.5°/min. The local structure of the samples was investigated by the WiTec Alpha 300 AR Raman microscopy system (objective lens, blue laser, Ulm, Germany). The morphology and chemical composition of the samples were studied using a scanning electron microscope Phenom ProX Desktop (SEM) (Waltham, MA, USA) integrated with energy-dispersive X-ray diffraction (EDS) detector.
The Perkin Elmer Pyris 1 TGA thermogravimetric analyzer (London, UK) was used for the investigation of thermal behavior of the hydrated phase. The heating of initially hydrated samples was made at the temperature range of 40–1100 °C with speed of 10 °C/min under a flow of dry Ar. The hydrated samples were obtained at slow cooling from 1100 to 150 °C (1 °C/min) under a flow of wet Ar (99.999% purity, pH2O = 2 × 10−2 atm). Ar atmosphere was used to avoid any carbonization of the samples.
The electrical conductivity was measured on the pressed cylindrical pellets using impedance spectroscopy method. The impedance spectrometer Z-1000P (Elins, RF, Uppsala, Sweden) with the frequency range of 1–106 Hz was used. The dependencies of conductivities vs. temperature were obtained in the temperature range of 200–1000 °C (step 10–20 °C, 1°/min cooling rate). All electrochemical investigations were performed under dry and wet air or Ar. The dry gas was produced by circulating the gas through P2O5 (pH2O = 3.5 × 10−5 atm). The wet gas was obtained by bubbling the gas at room temperature first through distilled water and then through saturated solution of KBr (pH2O = 2 × 10−2 atm). The humidity of the gas was controlled by a Honeywell HIH-3610 H2O sensor. The dependencies of conductivities vs. partial oxygen pressures pO2 were obtained by using the electrochemical method for producing different pO2 with oxygen pump (and sensor) from Y-stabilized ZrO2 ceramic. The values of the resistance were recorded after 3–5 h of equilibrium.

3. Results

3.1. Material Characterization

The powder samples BaM2In2O7 (M = La, Nd) were investigated using X-ray diffraction analysis. The both compositions were single phase, and they are indexed in the tetragonal symmetry (space group P42/mnm) (Figure 2a). The lattice and structural parameters were in good agreement with previously reported by Titov et al. and Raveau et al. data [27,28] (Table 1).
The morphology of the samples was investigated using scanning electron microscopy (SEM) method. The size of the grains was ~5–10 μm; the agglomerates with the size up to 30–50 μm were also found (Figure 2c). Figure 2d represents the image of the powder sample BaNd2In2O7. The chemical composition was checked via the energy-dispersive (EDS) analysis. The good agreement between theoretical and experimental values of chemical composition was proved by the energy-dispersive (EDS) analysis performed on the polished cleavages of the ceramic samples (Table 2).

3.2. Oxygen-Ionic Conductivity

The conductivity measurements were made using the impedance spectroscopy method. As an example, the typical Nyqiust plot is presented in Figure 3. The experimental data are showed by the blue symbols, and the fitting of the spectra (ZView software) is represented by the red line. According to the spectra, two different electrochemical processes can be defined. The semicircle started from the zero coordinates corresponds to the bulk resistance with the capacity ~10−12 F. The small semicircle in the low-frequency region is characterized by the capacity ~10−10 F and it corresponds to the resistance of the grain boundaries. It should be noted what the Nyqiust plots for the monolayer composition BaNdInO4 described by Yang et al. [33] was characterized by the same shape. For the calculation of electrical conductivity, the bulk resistance values were used.
The temperature dependencies of conductivity obtained under dry condition (pH2O = 3.5 × 10−5 atm) are presented in Figure 4a. As can be seen, the conductivity values for the composition BaNd2In2O7 are higher than for the BaLa2In2O7 [17]. The increase in the conductivity is about one order of magnitude under dry air condition. In the dry Ar, this difference is up to 1.2 order of magnitude. The dependencies of conductivity vs. oxygen partial pressure under dry condition are presented in Figure 4b. The right part of the curves (in the pO2 range of 10−5–0.21 atm) had a positive slope which corresponded to the electronic conductivity (p-type). The conductivity values are independent on the oxygen partial pressure below 10−5 atm, and this region belongs to the electrolytic area. It should be noted, that conductivity values obtained from “σ–1/T” dependencies (orange symbols, Figure 4b) well correlated with the values obtained from “σ–pO2” dependencies (blue symbols, Figure 4b). Based on this, the conductivity values obtained under dry Ar can be considered as the oxygen-ionic conductivity values. Consequently, the oxygen-ionic transport number t O 2 can be calculated as:
t O 2 = σ O 2 σ tot = σ Ar σ air
The oxygen-ionic transport numbers for the compounds BaLa2In2O7 and BaNd2In2O7 were around 20% and 50% correspondingly, i.e., the share of oxygen transport was bigger for the Nd-containing sample. For the explanation of this fact, the nature of oxygen transport should be considered.
The layered RP-structure BaM2In2O7 with n = 2 consists of alternating salt layers and perovskite blocks where two octahedra [InO6] layers are connected to each other by vertices (Figure 2b). The barium atoms are located in the perovskite blocks in the space between [InO6] octahedra and they have 12 coordination numbers. The atoms of rare earth elements (La and Nd) are placed in the salt layers and they characterized by a coordination number of 9. The increase in the coordination number of La or Nd atoms up to 12 is theoretically possible, and this process can be described in terms of anti-Frenkel disordering:
O O x V o + O i
where O i is the oxygen atom in the interstitial position; V O is the oxygen vacancy. In this case, the coordination number of La/Nd atoms increase, and the coordination number of In atoms decrease. Returning to the experimental data, we can say, that in the row BaLa2In2O7–BaNd2In2O7 not only total conductivity increases, but the share of oxygen-ionic transport increases also. Consequently, the different nature of rare earth elements (La or Nd) in the crystal lattice of layered perovskite BaM2In2O7 leads to the changes in the degree of disordering and the change in the local structure.
The local structure of compounds BaLa2In2O7 and BaNd2In2O7 was investigated using Raman spectroscopy. The results of the deconvolution of obtained Raman spectra are presented in Figure 5. The region of low wavenumbers (120–200 cm−1) is represented by the bending and stretching vibrations of polyhedral, containing alkali earth and rare earth metals [40,41,42,43], i.e., [BaO12], [LaO9], and [NdO9]. In this region, the modes ν1, ν2, ν3, and ν4 are observed (Table 3). They can be ascribed to the M–O stretching and O–M–O bending vibrations where M is the Ba, La, or Nd. For the composition BaNd2In2O7 the red shift (i.e., shift towards lower wavenumbers) is observed. In general, the red shift indicates the increase in the part of bond lengths M–O. However, due to the presence of several cations in the A- and A′-sublattices of AA 2 B 2 O 7 , the correct interpretation of this shift is difficult.
The region of mid and high wavenumbers (higher than 200 cm−1) contains the tilting/bending and stretching vibrations of In-contained polyhedra. The band ν5, ν6, ν7, ν8, and ν9 can be described as the tilting/bending vibration of polyhedra [InO6]. These five bands are well visible in the spectrum of BaLa2In2O7, while the spectrum of BaNd2In2O7 contains two (ν6 and ν8) bands only. It can be said, that ν6 and ν8 bands in the spectrum of BaNd2In2O7 degenerate to the ν5, ν6, ν7, ν8, and ν9 bands in the spectrum of BaLa2In2O7 correspondingly. Based on this, it can be assumed that tilting of In-containing octahedra decreases in the row BaLa2In2O7–BaNd2In2O7. According to the structural data reported by Titov et al. [27], the degree of deformation of [InO6] polyhedra decreases and the In–O–In angle is approaching to 180° from BaLa2In2O7 to BaNd2In2O7. The band ν14 registered in the spectrum of BaLa2In2O7 can be ascribed to the repulsion between the Ba2+/La3+(Nd3+) ions and apical oxygen ions in compressed In-contained polyhedra [44]. The decrease in the tilting of In-contained polyhedra in the row BaLa2In2O7–BaNd2In2O7 can lead to the disappearance of this band in the spectrum of BaNd2In2O7.
The stretching vibrations of In-contained polyhedra locate in the higher wavenumbers. Based on the analysis of Raman spectra of monolayer RP-composition BaLaInO4 [40,41], the stretching vibrations of In–O bonds should be appeared around 400 cm−1. At the same time, the comparable analysis of Raman spectra of RP homologous series Srn+1TinO3n+1 [45] and Srn+1RunO3n+1 [46] showed that the signal corresponded to the M–O stretching vibrations for monolayer compositions transforms into two signals with lower and higher wavenumbers in the spectra of double-layered RP-compositions. Thus, the bands ν10, ν11, and ν12 can be attributed to stretching vibrations of In–O bonds. As can be seen, the blue shift (i.e., shift towards higher wavenumbers) of ν10 and ν11 bands is observed for BaNd2In2O7 composition compared with BaLa2In2O7 composition. This shift indicates the decrease in some In–O bond lengths.
The local distortion of the crystal lattice due to anti-Frenkel disordering (Equation (2)) can lead to the formation of oxygen vacancies in the In-contained polyhedra, and the polyhedra with lower coordination number [InO6−x] appear in the structure. At the same time, the formation of oxygen interstitial leads to the increase in the coordination number of some cations. Most reasonable candidate for this is the rare earth element because both La and Nd have a coordination number of 9, and they can increase it up to 12 theoretically. Thus, the formation of polyhedra [La(Nd)O9+x] is possible. It is obvious that the decrease in the coordination number of metal leads to the decrease in some part of metal-oxygen bond length, and the signal with higher wavenumber should appear in the spectrum, what was observed for the composition BaLaInO4 [40,41]. We can assume that the signal ν13 in the spectra of BaNd2In2O7 and BaLa2In2O7 can be attributed to the motion of In-contained polyhedra with a lower coordination number.
The blue shift of ν13 band for the BaNd2In2O7 indicates that the decrease in the part of In–O bond lengths is more pronounced. Thus, we can assume that for the Nd-contained compound the formation of oxygen vacancies due to local distortion is more favorable in comparison with BaLa2In2O7 composition. This result is in good agreement with the crystallographic point of view, because the lower coordination number is more favorable for the ion with smaller ionic radius ( r Nd 3 + = 1.163 Å, r La 3 + = 1.216 Å [47]). In this way, the concentration of oxygen defects (oxygen vacancies and oxygen interstitial) should be higher in the composition BaNd2In2O7 compared with the composition BaLa2In2O7, and the share of oxygen-ionic conductivity should be higher also. Returning to obtained experimental data, we can say, that the results of electrical measurements are well correlated with the Raman spectroscopy results. The electrical conductivity and the share of oxygen-ionic transport increased in the row BaLa2In2O7–BaNd2In2O7.

3.3. Protonic Conductivity

The possibility of water uptake for the compounds with PR-structure is provided by the opportunity of the cations increase their coordination numbers. For the composition BaM2In2O7 the hydration may be due to increase in the coordination number of atoms M and the formation of polyhedra [MO8(OH)2]. This process can be described by the quasi-chemical equation as:
H 2 O + O o x ( OH ) o + ( OH ) i
where ( OH ) o is the hydroxyl group in the regular oxygen position; ( OH ) i is the hydroxyl group located in the interlayer space. However, in the case of the presence oxygen vacancies in the structure, the water uptake can be described also as:
V o + H 2 O + O o x OH o + OH V o
It is obvious that both of these processes are happened at the same time for the compositions with local distortion of the crystal lattice. Based on this, we can assume that in the structure of hydrated RP-compositions at least three different crystallographic positions of oxygen-hydrogen groups exist. Protons can be localized the structural oxygen atoms ( OH ) o and part of the oxygen-hydrogen groups may exist in the oxygen vacancies OH V o in the perovskite blocks and in the salt block in the interstitial oxygen positions ( OH ) i .
Figure 6 represents the thermogravimetric (TG) data of hydrated compositions BaLa2In2O7 and BaNd2In2O7. As we can see, three steps on the TG-curves for both compositions are observed. These results can be indirect evidence of the presence of oxygen vacancies in the structure of compositions BaM2In2O7. The amount of water uptake is almost the same for both compositions and it is around 0.15–0.17 mol H2O per formula BaM2In2O7 unit. Consequently, the difference between protonic conductivity values for the compositions BaLa2In2O7 and BaNd2In2O7 must be provided by the difference between protonic mobility only:
σ i = z i · e · μ i · c i
The temperature dependencies of conductivity obtained under wet condition are presented in Figure 7a (pH2O = 2 × 10−2 atm). The conductivity values for the composition BaNd2In2O7 are higher compared with the values for the composition BaLa2In2O7. In the region of high temperatures (T > 700 °C), where protonic transport is negligible, the conductivity values obtained under wet Ar are significantly lower than under wet air. In the region of low temperatures (T < 450 °C), the conductivity values obtained under wet air and wet Ar are almost the same. In other words, the conductivity values obtained under wet air below 450 °C can be considered as the values of ionic conductivity. This statement is proved by the data obtained from the “σ–pO2” dependencies. As can be seen in Figure 8, the conductivity values obtained under wet condition and low temperatures (T < 450 °C) are independent of the oxygen partial pressure even under oxidizing conditions (pO2 = 10−5–0.21 atm).
The effect of changes in the water partial pressure on the ionic conductivity is well visible in Figure 7b, where temperature dependencies of conductivity under dry and wet Ar are presented. As can be seen, the decrease in the temperature leads to the increase in the conductivity values obtained under wet Ar due to formation of the protonic charge carriers in the structure, which are more mobile in comparison with oxygen ions.
The protonic conductivity values were calculated as difference between the conductivity values in wet and dry Ar. The temperature dependencies of protonic conductivities for the compositions BaLa2In2O7 and BaNd2In2O7 are presented in Figure 9.
As can be seen, the values for BaNd2In2O7 are higher than for BaLa2In2O7 about one order of magnitude. The increase in the protonic conductivity values is well correlated with the decrease in the values of activation energy from 0.63 eV for BaLa2In2O7 to 0.5 eV for BaNd2In2O7. Returning to the results of TG-measurements, which showed closeness of protons concentrations for both compositions, we can say, that the increase in the protonic conductivity values for the composition BaNd2In2O7 is due to the increase in the protons mobility. The calculation of protonic transport numbers tH according to the formula:
t H = σ H + σ w e t   a i r
This showed, that for both compositions BaLa2In2O7 and BaNd2In2O7 with decreasing temperature the increase in tH occurs and the values reach up to 95% at the temperatures below 350 °C. Thus, these complex oxides can be considered as promising matrix compositions for creation of novel high-conductive protonic electrolytes with PR structure.

4. Conclusions

In this paper, the complex oxides BaLa2In2O7 and BaNd2In2O7 were synthesized, and the local structure, water uptake, and the electrical conductivity were investigated. It was shown that the values of electrical conductivity of double-layered complex oxides with PR structure BaM2In2O7 (M = La, Nd) are strongly depend on the nature of rare earth metal in the cationic sublattice. The change in the ionic radius of the element leads not only to the change in the unit cell parameters but also to the change in the local structure. It was shown that the composition BaNd2In2O7 is characterized by the higher values of electrical conductivity and the higher share of oxygen-ionic conductivity under dry air condition compared to the composition BaLa2In2O7. The most likely reason for this is an increase in local distortions in the crystal lattice and the formation of a greater number of oxygen defects. Under wet air conditions both compositions are nearly pure protonic conductors below 350 °C.

Author Contributions

Conceptualization, I.A. and N.T.; methodology, I.A., D.M. and N.T.; investigation, A.G., K.B., E.A. and A.E.; data curation, N.T., K.B., E.A., A.E. and A.G.; writing—original draft preparation, N.T.; writing—review and editing, N.T., I.A. and D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was performed according to the budgetary plan of the Institute of High Temperature Electrochemistry and funded by the Budget of Russian Federation.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ruddlesden, S.N.; Popper, P. The compound Sr3Ti2O7 and its structure. Acta Cryst. 1958, 11, 54–55. [Google Scholar] [CrossRef] [Green Version]
  2. Shimizu, K.; Tsuji, Y.; Hatamachi, T.; Toda, K.; Kodama, T.; Sato, M.; Kitayama, Y. Photocatalytic water splitting on hydrated layered perovskite tantalate A2SrTa2O7·nH2O (A = H, K, and Rb). Phys. Chem. Chem. Phys. 2004, 6, 1064–1069. [Google Scholar] [CrossRef]
  3. Liang, Z.; Tang, K.; Shao, Q.; Lia, G.; Zeng, S.; Zheng, H. Synthesis, crystal structure, and photocatalytic activity of a new two-layer Ruddlesden–Popper phase, Li2CaTa2O7. J. Solid State Chem. 2008, 181, 964–970. [Google Scholar] [CrossRef]
  4. Wang, Y.; Wang, C.; Wang, L.; Hao, Q.; Zhu, X.; Chen, X.; Tang, K. Preparation of interlayer surface tailored protonated double-layered perovskite H2CaTa2O7 with n-alcohols, and their photocatalytic activity. RSC Adv. 2014, 4, 4047–4054. [Google Scholar] [CrossRef]
  5. Wang, C.; Jin, Y.; Lv, Y.; Ju, G.; Chen, L.; Li, X.; Hu, Y. A bifunctional phosphor Sr3Sn2O7:Eu3+: Red luminescence and photochromism properties. J. Lumin. 2017, 192, 337–342. [Google Scholar] [CrossRef]
  6. Zhang, L.; Sun, B.; Liu, Q.; Ding, N.; Yang, H.; Wang, L.; Zhang, Q. Novel layered perovskite Sr3Ti2O7:Eu3+ phosphor with high-efficiency luminescence enhanced by charge compensation. J. Alloys Compd. 2016, 657, 27–31. [Google Scholar] [CrossRef]
  7. Battle, P.D.; Rosseinsky, M.J. Synthesis, structure, and magnetic properties of n=2 Ruddlesden–Popper manganates. Curr. Opin. Solid State Mater. Sci. 1999, 4, 163–170. [Google Scholar] [CrossRef]
  8. Ling, C.D.; Millburn, J.E.; Mitchell, J.F.; Argyriou, D.N.; Linton, J. Interplay of spin and orbital ordering in the layered colossal magnetoresistance manganite La2–2xSr1+2xMn2O7 (0.5 ≤ x ≤ 1.0). Phys. Rev. B 2000, 62, 15096–15111. [Google Scholar] [CrossRef] [Green Version]
  9. Medvedeva, J.E.; Anisimov, V.I.; Mryasov, O.N.; Freeman, A.J. The role of Coulomb correlation in magnetic and transport properties of doped manganites: La0.5Sr0.5MnO3 and LaSr2Mn2O7. J. Phys. Condens. Matter 2002, 14, 4533–4542. [Google Scholar] [CrossRef] [Green Version]
  10. Taşarkuyu, E.; Coşkun, A.; Irmak, A.E.; Aktürk, S.; Ünlüa, G.; Samancıoğlu, Y.; Yücel, A.; Sarıkürkçü, C.; Aksoy, S.; Acet, M. Effect of high temperature sintering on the structural and the magnetic properties of La1.4Ca1.6Mn2O7. J. Alloys Compd. 2011, 509, 3717–3722. [Google Scholar] [CrossRef]
  11. Dudric, R.; Goga, F.; Neumann, M.; Mican, S.; Tetean, R. Magnetic properties and magnetocaloric effect in La1.4–xCexCa1.6Mn2O7 perovskites synthesized by sol–gel method. J. Mater. Sci. 2012, 47, 3125–3130. [Google Scholar] [CrossRef]
  12. Denbri, F.; Mahamdioua, N.; Meriche, F.; Altintas, S.P.; Terzioglu, C. Investigation of magneto-transport properties of the codoped La1.6-xPrxCa1.4-xBaxMn2O7 (x = 0.2 and 0.4) double-layered manganite. J. Mater. Sci. Mater. Electron. 2021, 32, 18808–18824. [Google Scholar] [CrossRef]
  13. Boulahya, K.; Muñoz-Gil, D.; Gómez-Herrero, A.; Azcondo, M.T.; Amador, U. Eu2SrCo1.5Fe0.5O7 a new promising Ruddlesden–Popper member as cathode component for intermediate temperature solid oxide fuel cells. J. Mater. Chem. A 2019, 7, 5601–5611. [Google Scholar] [CrossRef]
  14. Wang, Q.; Hou, J.; Fan, Y.; Xi, X.; Li, J.; Lu, Y.; Huo, G.; Shao, L.; Fu, X.; Luo, J. Pr2BaNiMnO7-δ double-layered Ruddlesden–Popper perovskite oxides as efficient cathode electrocatalysts for low temperature proton conducting solid oxide fuel cells. J. Mater. Chem. A 2020, 8, 7704–7712. [Google Scholar] [CrossRef]
  15. Muñoz Gil, D.; Boulahya, K.; Santoyo, M.S.; Azcondo, M.T.; Amador, U. Superior Performance as Cathode Material for Intermediate-Temperature Solid Oxide Fuel Cells of the Ruddlesden−Popper n = 2 Member Eu2SrCo0.50Fe1.50O7−δ with Low Cobalt Content. Inorg. Chem. 2021, 60, 3094–3105. [Google Scholar] [CrossRef]
  16. Kananke-Gamage, C.C.W.; Ramezanipou, F. Variation of the electrocatalytic activity of isostructural oxides Sr2LaFeMnO7 and Sr2LaCoMnO7 for hydrogen and oxygen-evolution reactions. Dalton Trans. 2021, 50, 14196–14206. [Google Scholar] [CrossRef]
  17. Tarasova, N.; Galisheva, A.; Animitsa, I.; Korona, D.; Kreimesh, H.; Fedorova, I. Protonic transport in layered perovskites BaLanInnO3n+1 (n = 1, 2) with Ruddlesden-Popper structure. Appl. Sci. 2022, 12, 4082. [Google Scholar] [CrossRef]
  18. Kim, S.; Kim, G.; Manthiram, A. A review on infiltration techniques for energy conversion and storage devices: From fundamentals to applications. Sustain. Energy Fuels 2021, 5, 5024–503721. [Google Scholar] [CrossRef]
  19. Klyndyuk, A.I.; Chizhova, E.A.; Kharytonau, D.S.; Medvedev, D.A. Layered oxygen-deficient double perovskites as promising cathode materials for solid oxide fuel cells. Materials 2022, 15, 141. [Google Scholar] [CrossRef]
  20. Hanif, M.B.; Motola, M.; Qayyum, S.; Rauf, S.; Khalid, A.; Li, C.-J.; Li, C.-X. Recent advancements, doping strategies and the future perspective of perovskite-based solid oxide fuel cells for energy conversion. Chem. Eng. J. 2022, 42815, 132603. [Google Scholar] [CrossRef]
  21. Shim, J.H. Ceramics breakthrough. Nat. Energy 2018, 3, 168–169. [Google Scholar] [CrossRef]
  22. Meng, Y.; Gao, J.; Zhao, Z.; Amoroso, J.; Tong, J.; Brinkman, K.S. Review: Recent progress in low-temperature proton-conducting ceramics. J. Mater. Sci. 2019, 54, 9291–9312. [Google Scholar] [CrossRef] [Green Version]
  23. Kim, J.; Sengodan, S.; Kim, S.; Kwon, O.; Bud, Y.; Kim, G. Proton conducting oxides: A review of materials and applications for renewable energy conversion and storage. Renew. Sustain. Energy Rev. 2019, 109, 606–618. [Google Scholar] [CrossRef]
  24. Medvedev, D.A. Current drawbacks of proton-conducting ceramic materials: How to overcome them for real electrochemical purposes. Curr. Opin. Green Sustain. Chem. 2021, 32, 100549. [Google Scholar] [CrossRef]
  25. Bello, I.T.; Zhai, S.; He, Q.; Cheng, C.; Dai, Y.; Chen, B.; Zhang, Y.; Ni, M. Materials development and prospective for protonic ceramic fuel cells. Int. J. Energy Res. 2022, 46, 2212–2240. [Google Scholar] [CrossRef]
  26. Irvine, J. Roadmap on inorganic perovskites for energy applications. J. Phys. Energy 2021, 3, 031502. [Google Scholar] [CrossRef]
  27. Titov, Y.A.; Belyavina, N.M.; Markiv, V.Y.; Slobodyanik, M.S.; Krayevska, Y.A.; Yaschuk, V.P. Synthesis and crystal structure of BaLn2In2O7. Rep. Natl. Acad. Sci. Ukr. 2010, 1, 148–153. [Google Scholar]
  28. Caldes, M.; Michel, C.; Rouillon, T.; Hervieu, M.; Raveau, B. Novel indates Ln2BaIn2O7, n = 2 members of the Ruddlesden–Popper family (Ln = La, Nd). J. Mater. Chem. 2002, 12, 473–476. [Google Scholar] [CrossRef]
  29. Fujii, K.; Esaki, Y.; Omoto, K.; Yashima, M.; Hoshikawa, A.; Ishigaki, T.; Hester, J.R. New Perovskite-Related Structure Family of Oxide-Ion Conducting Materials NdBaInO4. Chem. Mater. 2014, 26, 2488–2491. [Google Scholar] [CrossRef]
  30. Fujii, K.; Shiraiwa, M.; Esaki, Y.; Yashima, M.; Kim, S.J.; Lee, S. Improved oxide-ion conductivity of NdBaInO4 by Sr doping. J. Mater. Chem. A 2015, 3, 11985. [Google Scholar] [CrossRef] [Green Version]
  31. Ishihara, T.; Yan, Y.; Sakai, T.; Ida, S. Oxide ion conductivity in doped NdBaInO4. Solid State Ionics 2016, 288, 262–265. [Google Scholar] [CrossRef]
  32. Yang, X.; Liu, S.; Lu, F.; Xu, J.; Kuang, X. Acceptor Doping and Oxygen Vacancy Migration in Layered Perovskite NdBaInO4-Based Mixed Conductors. J. Phys. Chem. C 2016, 120, 6416–6426. [Google Scholar] [CrossRef]
  33. Fijii, K.; Yashima, M. Discovery and development of BaNdInO4—A brief review. J. Ceram. Soc. Jpn. 2018, 126, 852–859. [Google Scholar] [CrossRef] [Green Version]
  34. Zhou, Y.; Shiraiwa, M.; Nagao, M.; Fujii, K.; Tanaka, I.; Yashima, M.; Baque, L.; Basbus, J.F.; Mogni, L.V.; Skinner, S.J. Protonic Conduction in the BaNdInO4 Structure Achieved by Acceptor Doping. Chem. Mater. 2021, 33, 2139–2146. [Google Scholar] [CrossRef]
  35. Tarasova, N.; Galisheva, A.; Animitsa, I. Ba2+/Ti4+- co-doped layered perovskite BaLaInO4: The structure and ionic (O2−, H+) conductivity. Int. J. Hydrogen Energy 2021, 46, 16868–16877. [Google Scholar] [CrossRef]
  36. Tarasova, N.A.; Galisheva, A.O.; Animitsa, I.E.; Lebedeva, E.L. Oxygen-Ion and Proton Transport in Sc-Doped Layered Perovskite BaLaInO4. Russ. J. Electrochem. 2021, 57, 1008–1014. [Google Scholar] [CrossRef]
  37. Tarasova, N.A.; Galisheva, A.O.; Animitsa, I.E.; Dmitrieva, A.A. The Effect of Donor Doping on the Ionic (O2−, H+) Transport in Novel Complex Oxides BaLaIn1–xNbxO4+x with the Ruddlesden–Popper Structure. Russ. J. Electrochem. 2021, 57, 962–969. [Google Scholar] [CrossRef]
  38. Tarasova, N.; Animitsa, I.; Galisheva, A. Effect of acceptor and donor doping on the state of protons in block-layered structures based on BaLaInO4. Solid State Commun. 2021, 323, 14093. [Google Scholar] [CrossRef]
  39. Tarasova, N.; Animitsa, I.; Galisheva, A. Spectroscopic and transport properties of Ba- and Ti-doped BaLaInO4. J. Raman Spec. 2021, 52, 980–987. [Google Scholar] [CrossRef]
  40. Tarasova, N.; Animitsa, I.; Galisheva, A. Effect of doping on the local structure of new block-layered proton conductors based on BaLaInO4. J. Raman Spec. 2021, 51, 2290–2297. [Google Scholar] [CrossRef]
  41. Tarasova, N.; Animitsa, I. Materials AIILnInO4 with Ruddlesden-Popper structure for electrochemical applications: Relationship between ion (oxygen-ion, proton) conductivity, water uptake and structural changes. Materials 2022, 15, 114. [Google Scholar] [CrossRef] [PubMed]
  42. Scherban, T.; Villeneuve, R.; Abello, L.; Lucazeau, G. Raman scattering study of acceptor-doped BaCeO3. Solid State Ion. 1993, 61, 93–98. [Google Scholar] [CrossRef]
  43. Chemarin, C.; Rosman, N.; Pagnier, T.; Lucazeau, G. High-Pressure Raman Study of Mixed Perovskites BaCexZr1-xO3 (0≤x≤1). J. Solid State Chem. 2000, 149, 298–307. [Google Scholar] [CrossRef]
  44. Paul, B.; Chatterjee, S.; Gop, S.; Roy, A.; Grover, V.; Shukla, R.; Tyagi, A.K. Evolution of lattice dynamics in ferroelectric hexagonal REInO3 (RE = Ho, Dy, Tb, Gd, Eu, Sm) perovskites. Mater. Res. Express 2016, 3, 075703. [Google Scholar] [CrossRef]
  45. Kamba, S.; SamoukPina, F.; Kadlec, F.; Pokorny, J.; Petzelt, J.; Reaney, I.M.; Wise, P.L. Composition dependence of the lattice vibrations in Srn+1TinO3n+1 Ruddlesden–Popper homologous series. J. Eur. Ceram. Soc. 2003, 23, 2639–2645. [Google Scholar] [CrossRef]
  46. Iliev, M.N.; Popov, V.N.; Litvinchuk, A.P.; Abrashev, M.V.; Backstrom, J.; Sun, Y.Y.; Mena, R.L.; Chu, C.W. Comparative Raman studies of Sr2RuO4; Sr3Ru2O7 and Sr4Ru3O10. Physica B 2005, 358, 138–152. [Google Scholar] [CrossRef] [Green Version]
  47. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
Figure 1. The overview of materials with a double-layered Ruddlesden–Popper structure.
Figure 1. The overview of materials with a double-layered Ruddlesden–Popper structure.
Materials 15 03488 g001
Figure 2. Materials characterization of the BaNd2In2O7 compound: XRD patterns (a), image of double-layered RP crystal structure (b), SEM-image (c), and image of powder sample (d).
Figure 2. Materials characterization of the BaNd2In2O7 compound: XRD patterns (a), image of double-layered RP crystal structure (b), SEM-image (c), and image of powder sample (d).
Materials 15 03488 g002
Figure 3. The Nyquist-plot for the BaNd2In2O7 ceramic material obtained at 350 °C under dry air.
Figure 3. The Nyquist-plot for the BaNd2In2O7 ceramic material obtained at 350 °C under dry air.
Materials 15 03488 g003
Figure 4. Temperature dependencies of conductivity for BaNd2In2O7 (blue symbols) and BaLa2In2O7 (black symbols) [17] obtained for dry Ar (open symbols) and dry air (filled symbols) (a); jxygen partial pressure dependencies of the total conductivity values for the BaNd2In2O7 sample at dry conditions (blue symbols) and conductivity values from σ—103/T dependencies in dry air and Ar (orange symbols) (b).
Figure 4. Temperature dependencies of conductivity for BaNd2In2O7 (blue symbols) and BaLa2In2O7 (black symbols) [17] obtained for dry Ar (open symbols) and dry air (filled symbols) (a); jxygen partial pressure dependencies of the total conductivity values for the BaNd2In2O7 sample at dry conditions (blue symbols) and conductivity values from σ—103/T dependencies in dry air and Ar (orange symbols) (b).
Materials 15 03488 g004
Figure 5. Raman spectra of the BaLa2In2O7 (1) and BaNd2In2O7 (2) compounds.
Figure 5. Raman spectra of the BaLa2In2O7 (1) and BaNd2In2O7 (2) compounds.
Materials 15 03488 g005
Figure 6. Thermogravimetry data for the BaLa2In2O7 (a) and BaNd2In2O7 (b) compounds obtained under wet (1) and dry (2) Ar.
Figure 6. Thermogravimetry data for the BaLa2In2O7 (a) and BaNd2In2O7 (b) compounds obtained under wet (1) and dry (2) Ar.
Materials 15 03488 g006
Figure 7. Temperature dependencies of conductivity for BaNd2In2O7 (blue symbols) and BaLa2In2O7 (black symbols) [17] obtained at wet Ar (open symbols) and wet air (filled symbols) (a); wet Ar (open symbols) and dry Ar (filled symbols) (b).
Figure 7. Temperature dependencies of conductivity for BaNd2In2O7 (blue symbols) and BaLa2In2O7 (black symbols) [17] obtained at wet Ar (open symbols) and wet air (filled symbols) (a); wet Ar (open symbols) and dry Ar (filled symbols) (b).
Materials 15 03488 g007
Figure 8. Oxygen partial pressure dependencies of the total conductivity for the BaNd2In2O7 sample at dry (filled symbols) and wet (open symbols) conditions and conductivity values from σ–103/T dependencies (orange symbols).
Figure 8. Oxygen partial pressure dependencies of the total conductivity for the BaNd2In2O7 sample at dry (filled symbols) and wet (open symbols) conditions and conductivity values from σ–103/T dependencies (orange symbols).
Materials 15 03488 g008
Figure 9. Temperature dependencies of protonic conductivity for BaNd2In2O7 (blue line) and BaLa2In2O7 (black line).
Figure 9. Temperature dependencies of protonic conductivity for BaNd2In2O7 (blue line) and BaLa2In2O7 (black line).
Materials 15 03488 g009
Table 1. Unit cell parameters and volume of the BaNd2In2O7 and BaLa2In2O7 compounds.
Table 1. Unit cell parameters and volume of the BaNd2In2O7 and BaLa2In2O7 compounds.
CompositionUnit Cell ParametersThis WorkTitov et al. [27]Raveau et al. [28]
BaNd2In2O7a, Å5.8916(9)5.8969(8)5.8940(3)
c, Å20.469(0)20.439(3)20.467(1)
V, Å3710.52710.76711.02
BaLa2In2O7a, Å5.914(9)5.915(2)5.9141(3)
c, Å20.846(5)20.861(1)20.831(2)
V, Å3729.34729.92729.73
Table 2. The average metal ratios determined by EDS analysis for the BaNd2In2O7 and BaLa2In2O7 compounds (theoretical values are given in the brackets).
Table 2. The average metal ratios determined by EDS analysis for the BaNd2In2O7 and BaLa2In2O7 compounds (theoretical values are given in the brackets).
ElementContent of the Elements, Atomic %
BaNd2In2O7BaLa2In2O7
Ba8.81 (8.33)8.42 (8.33)
La/Nd16.23 (16.67)16.37 (16.67)
In16.95 (16.67)16.71 (16.67)
O58.01 (58.33)58.50 (58.33)
Table 3. Wavenumbers (cm−1) of Raman bands for the BaNd2In2O7 and BaLa2In2O7 compounds.
Table 3. Wavenumbers (cm−1) of Raman bands for the BaNd2In2O7 and BaLa2In2O7 compounds.
No of BandBaLa2In2O7BaNd2In2O7
1120118
2140133
3161157
4203181
5237-
6260250
7279-
8316319
9336-
10358370
11430453
12445-
13545660
14800-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tarasova, N.; Galisheva, A.; Animitsa, I.; Belova, K.; Egorova, A.; Abakumova, E.; Medvedev, D. Layered Perovskites BaM2In2O7 (M = La, Nd): From the Structure to the Ionic (O2−, H+) Conductivity. Materials 2022, 15, 3488. https://doi.org/10.3390/ma15103488

AMA Style

Tarasova N, Galisheva A, Animitsa I, Belova K, Egorova A, Abakumova E, Medvedev D. Layered Perovskites BaM2In2O7 (M = La, Nd): From the Structure to the Ionic (O2−, H+) Conductivity. Materials. 2022; 15(10):3488. https://doi.org/10.3390/ma15103488

Chicago/Turabian Style

Tarasova, Nataliia, Anzhelika Galisheva, Irina Animitsa, Ksenia Belova, Anastasia Egorova, Ekaterina Abakumova, and Dmitry Medvedev. 2022. "Layered Perovskites BaM2In2O7 (M = La, Nd): From the Structure to the Ionic (O2−, H+) Conductivity" Materials 15, no. 10: 3488. https://doi.org/10.3390/ma15103488

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