Next Article in Journal
Aliphatic-Bridged Early Lanthanide Metal–Organic Frameworks: Topological Polymorphism and Excitation-Dependent Luminescence
Next Article in Special Issue
Oxygen-Ion and Proton Transport of Origin and Ca-Doped La2ZnNdO5.5 Materials
Previous Article in Journal
Advances of Biowaste-Derived Porous Carbon and Carbon–Manganese Dioxide Composite in Supercapacitors: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxygen Ion and Proton Transport in Alkali-Earth Doped Layered Perovskites Based on BaLa2In2O7

by
Nataliia Tarasova
1,2,*,
Anzhelika Bedarkova
1,2,
Irina Animitsa
1,2,
Ksenia Belova
1,2,
Ekaterina Abakumova
1,2,
Polina Cheremisina
1,2 and
Dmitry Medvedev
1,2
1
Institute of High Temperature Electrochemistry of the Ural Branch of the Russian Academy of Sciences, Akademicheskaya St., 620066 Yekaterinburg, Russia
2
Institute of Hydrogen Energy, Ural Federal University, Mira St. 19, 620075 Yekaterinburg, Russia
*
Author to whom correspondence should be addressed.
Inorganics 2022, 10(10), 161; https://doi.org/10.3390/inorganics10100161
Submission received: 14 September 2022 / Revised: 21 September 2022 / Accepted: 29 September 2022 / Published: 1 October 2022
(This article belongs to the Special Issue Layered Perovskites: Synthesis, Properties and Structures)

Abstract

:
Inorganic materials with layered perovskite structures have a wide range of physical and chemical properties. Layered perovskites based on BaLanInnO3n+1 (n = 1, 2) were recently investigated as protonic conductors. This work focused on the oxygen ion and proton transport (ionic conductivity and mobility) in alkali-earth (Sr2+, Ba2+)-doped layered perovskites based on BaLa2In2O7. It is shown that in the dry air conditions, the nature of conductivity is mixed oxygen–hole, despite the dopant nature. Doping leads to the increase in the conductivity values by up to ~1.5 orders of magnitude. The most proton-conductive BaLa1.7Ba0.3In2O6.85 and BaLa1.7Sr0.15In2O6.925 samples are characterized by the conductivity values 1.2·10−4 S/cm and 0.7·10−4 S/cm at 500 °C under wet air, respectively. The layered perovskites with Ruddlesden-Popper structure, containing two layers of perovskite blocks, are the prospective proton-conducting materials and further material science searches among this class of materials is relevant.

1. Introduction

The layered perovskite family includes such classes of perovskite-related structures as Ruddlesden-Popper (RP), Dion–Jacobson (DJ), and Aurivillius structures (Figure 1). Compounds with a Ruddlesden-Popper structure have the general formula An+1BnO3n+1, in which the rock salt layers AO alternate with perovskite blocks (ABO3)n. This type of structure was described for the first time for the compositions Sr2TiO4 [1] and Sr2Ti2O7 [2] by S.N. Ruddlesden and P. Popper in 1957–1958. Aurivillius phases are characterized by the formula An−1Bi2BnO3n+3, which describes alternation of perovskite layers An−1BnO3n+1 with layers of a fluorite structure, formed by bismuth and oxygen ions [Bi2O2]2+. This crystal structure was first described in 1949 by B. Aurivillius [3]. Dion–Jacobson phases can be described by the formula AAn−1BnO3n+1. The structure of these compounds includes perovskite blocks An−1BnO3n+1 separated by layers, in which only metal cations A are presented. The mixed niobates of alkali and alkaline earth metals with such type of structure were synthesized for the first time by M. Dion [4] and A.J. Jacobson [5,6] in the first half of the 1980s.
The materials with layered perovskite-related structures have many various applications due to their different physical–chemical properties. They are known as photocatalysts [7,8,9,10,11,12,13], and include materials for solar hydrogen production [14,15,16,17], ferroelectrics [18,19,20,21,22], and phosphors [22,23,24,25,26,27,28]. Some of RP and DJ materials are capable of water molecule intercalation into the interlayer space. Ion exchange leads to the formation of protonated forms of compositions, for example HxLn1−xTiO4∙nH2O [29,30,31,32] or H1−xLaxCa2−xNb3O10 [33,34]. Recently, the possibility of dissociative incorporation of water molecules into crystal lattice was proven for the RP compositions based on AA′BO4 (A = Ba, Sr, A′ = La, Nd, B = In, Sc) and AA′2B2O7 (A = Ba, A′ = La, Nd, B = In). The materials based on the monolayer perovskites BaNdInO4 [35,36,37,38,39,40], BaLaScO4 [41], SrLaInO4 [42,43,44,45,46], and BaLaInO4 [47,48,49,50,51,52] were investigated as protonic conductors in the temperature range of 300–500 °C. It is shown that doping of the cationic sublattices is a successful way to improve the oxygen ion and proton transport [53]. Two-layer perovskites BaLa2In2O7 [54] and BaNd2In2O7 [55] are also investigated as proton conductors. The acceptor doping of the lanthanum sublattice of BaLa2In2O7 allows for an increase in the ionic conductivity of the matrix composition [56]. In this work, the effect of acceptor dopant (Sr2+, Ba2+) concentration on the ionic transport (O2−, H+) of BaLa2−xMxIn2O7−0.5x solid solutions, including the effect on the oxygen and proton mobility, was found.

2. Results and Discussion

2.1. XRD and TG Investigations

The X-ray diffraction (XRD) analysis was applied to establish the homogeneity ranges of the solid solutions BaLa2−xMxIn2O7−0.5x (M = Sr, Ba). The single-phase compositions were obtained in the ranges 0 ≤ x ≤ 0.20 and 0 ≤ x ≤ 0.30 for the Sr- and Ba-doped solid solutions, respectively (tetragonal symmetry, P42/mnm space group). The introduction of ions with bigger ionic radius ( r La 3 + = 1.216 Å, r Sr 2 + = 1.31 Å, r Ba 2 + = 1.47 Å [57]) leads to the increase in the lattice parameters and unit cell volumes of doped compositions (Table 1 and Table 2). Figure 2a represents the example of a full-profile analysis for the BaLa1.8Ba0.2In2O6.9 composition.
The possibility for water uptake was investigated by the thermogravimetry (TG) method. The amounts of water uptake for each composition are presented in the Table 1 and Table 2. As can be seen, water uptake for all undoped and doped compositions is close, and it is in the range 0.15–0.22 mol water per formula unit. It should be added that these values are comparable with water uptake for acceptor-doped classic perovskites. The TG curve coupled with MS water signal for the hydrated composition BaLa1.8Ba0.2In2O6.9∙0.2H2O are presented in Figure 3.
The dissociative incorporation of water for the classic perovskites ABO3−δ is due to the interaction of oxygen vacancies V o (appearing by the acceptor doping) or V o × (own structural defects) with water molecules:
V o + H 2 O + O o × 2 OH o ,
V o × + H 2 O + 2 O o × 2 OH o + O i ,
where V o is the oxygen vacancy, O o × is the oxygen atom in the regular position,   OH o is the hydroxyl group in the oxygen sublattice, and O i is the oxygen atom in the interstitial position.
Accordingly, the amount of water uptake must increase with increase in the oxygen vacancy concentration in the crystal structure of a doped complex oxide:
2 MO   La 2 O 3 2 M La + 2 O o × + V o ,
where M La :Sr or Ba atoms in La sites; V o : oxygen vacancy; and O o × : oxygen atom in a regular position. However, the water uptake for the monolayer perovskites based on BaLaInO4 is much bigger than oxygen vacancy concentration, and it depends on the space size between perovskite blocks in the structure [53]. In this case, the dissociative intercalation of water can be described as the incorporation of hydroxyl groups into space between perovskite blocks:
H 2 O + O o × ( OH ) o + ( OH ) i ,
where ( OH ) o is the hydroxyl group in the regular oxygen position and ( OH ) i is the hydroxyl group located in the rock salt block. It is also shown that the water uptake for the two-layer BaLa2In2O7 composition is lower compared to the monolayer BaLaInO4 composition (0.17 vs. 0.62 mol per formula unit, respectively) [55]. In other words, monolayer and two-layer barium–lanthanum indates have different hydration relationships. The obtained results concerning water uptake for BaLa1−xMxIn2O7−x (M = Sr, Ba) also prove this. Obviously, the presence in the structure of two octahedral layers of perovskite blocks means less structural flexibility of the crystal lattice during hydration, and this factor is more significant than the increase in the lattice parameter and oxygen vacancies concentration during doping. In contrast to the previously described hydrated solid solutions based on BaLaInO4∙nH2O, the investigated hydrated solid solutions based on BaLa2−xMxIn2O7−0.5x∙nH2O do not change the symmetry; these solid solutions are described by the same P42/mnm space group as anhydrous BaLa2−xMxIn2O7−0.5x (M = Sr, Ba) phases also (Figure 2b).

2.2. Electrical Conductivity Investigations

The general view of the temperature dependencies of conductivities is the same for the all investigated compositions of BaLa2−xMxIn2O7−0.5x (M = Sr, Ba). As an example, the temperature dependencies of conductivity obtained under different oxygen and water partial pressure are presented in Figure 4 for the composition BaLa1.75Ba0.25In2O6.875.
As can be seen, the values obtained under low pO2 (~10−5 atm, Ar) and low pH2O (3.5·10−5 atm, dry atmosphere) conditions are lower than under dry air (pO2 = 0.21 atm, pH2O = 3.5·10−5 atm) conditions. The difference is about 0.5 order of magnitude over the whole investigated temperature range. The effect of high water partial pressure (pH2O = 2·10−2 atm, wet atmosphere) on the conductivity values is more significant at the temperatures below ~500 °C. In this temperature region (300–500 °C), conductivity values under wet conditions are higher that under dry conditions by about 0.4 order of magnitude under air and 0.8 order under Ar. In addition, the conductivity values under wet air and wet Ar are very close at T < 500 °C. Figure 5 represents the concentration dependencies of conductivities for both solid solutions BaLa2−xMxIn2O7−0.5x (M = Sr, Ba) at 500 °C under dry air. As can be seen, the maximum for the concentration dependencies of conductivities is observed for both solid solutions. The difference in the conductivity values under different oxygen partial pressure (air and Ar) indicates the mixed ionic–electronic nature of electrical conductivity (Figure 4). Consequently, the detailed analysis of the dependence of conductivity on oxygen partial pressure is necessary.
The σ−pO2 dependencies obtained under dry and wet conditions for the doped compositions are presented in Figure 6. All dependencies for Sr and Ba compositions have the same regularities, which are discussed below. It should be noted that the shape of σ−pO2 dependencies for undoped and doped compositions are also similar [54].
Under dry air and oxidizing conditions (pO2 > 10−4 atm), the nature of conductivity is mixed ionic–hole, and the process of the appearance of holes can be written by the following equation:
V o + 1 2 O 2 O o × + 2 h ,
where V o is the oxygen vacancy, O o × is the oxygen atom in the regular position, and h is the hole. The area of predominant oxygen ionic conductivity is observed at pO2, lower than 10−4 atm. The conductivity values obtained under the Ar atmosphere are localized in the electrolytic region (black open symbols in Figure 6), which allows us to consider the values obtained in dry argon as oxygen ion conductivity values.
The concentration dependencies of oxygen ion conductivity obtained at different temperatures are presented in Figure 7. As can be seen, these dependencies exhibit a similar shape to the dependencies of total (mixed ionic–electronic) conductivities (Figure 5). The maxima on the curves are recorded for the Sr-doped composition at x = 0.1 and for Ba-doped compositions at x = 0.15–0.25. In the other words, doping by both Sr2+ and Ba2+ ions leads to an increase in the oxygen ionic conductivity values. The possible reasons are the increase in the oxygen vacancies concentration and the expansion of space for ionic transport (increase in the lattice parameters and unit cell volumes) during doping. At the same time, the conductivity decreases in the area of high dopant concentrations. The most possible cause may be an association of point defects and the formation of the clusters:
M La + V o M La · V o   or
M La + M La · V o 2 M La · V o × ,
The observed decrease in the oxygen vacancy mobility at the high dopant concentrations (Figure 8) proves this assumption. It should be noted that the similar effect of decreasing the oxygen ion conductivity and mobility is observed for the Ba-doped solid solution based on monolayer BaLaInO4 composition [47].
The interaction of water molecules with the crystal lattice of layered perovskites based on BaLa2In2O7 can be expressed by Equation (4) or by the following reaction:
BaLa 2 In 2 O 7 + x 2 H 2 O BaLa 2 In 2 O 7 + x 2 OH x .
Consequently, the hydrated forms of complex oxides can be written, for example, as BaLa2In2O6.83(OH)0.34 (undoped composition). The effect of the water partial pressure on conductivity is more pronounced in the electrolytic region (Figure 6). The increase in the proton concentration during decreasing temperature leads to the decrease in the hole concentration and, consequently, in the hole conductivity:
h + 1 2 H 2 O + O o × 1 4 O 2 + OH o .
The predominance of protonic transport is achieved below ~450 °C. The good comparability between conductivity values obtained under wet Ar (black open signs in Figure 6) and values obtained from σ–pO2 dependencies at pO2 = 10−5 atm should be noted.
Proton conductivity values were calculated as the difference between the conductivity values in wet and dry Ar, and its concentration dependencies for the solid solutions BaLa2−xMxIn2O7−0.5x are presented in Figure 9. As can be seen, the proton conductivity increases with increasing dopant concentration. However, the most significant increase is observed in the area of “small” dopant concentrations (x ≤ 0.10 for Sr-doped and x ≤ 0.15 s for Ba-doped solid solutions). The further increase in the dopant concentration leads to the small increase in conductivity (Ba-doped solid solution) or results in no effect (Sr-doped solid solution).
The concentration dependencies of proton mobilities at 300 °C are presented in Figure 10. As can be seen, the increase in the proton mobility is observed in the area of “small” dopant concentration, and it can be associated with the increase in the oxygen ion mobility in this area (Figure 8). The further increase in the dopant concentration leads to the decrease in the proton concentration (Sr-doped solid solution) or to an insufficient increase (Ba-doped solid solution). Obviously, the cluster formation occurs in the wet atmosphere also:
M La + ( OH ) o ( M La ( OH ) o ) × ,
and this process is more significant for the compositions with smaller dopant size M La (that is, for the Sr-doped compositions). It should be noted that the same effect of decreasing the proton mobility under “high” dopant concentrations is observed for the Ba-doped solid solution based on the monolayer BaLaInO4 composition [47].
Thus, the acceptor doping of two-layer perovskite BaLa2In2O7 is a successful way for improving proton conductivity and obtaining novel high-conductive electrolytic materials. The doping by the Sr2+- and Ba2+- ions on the La3+ sublattice leads to the increase in the ion conductivity by up to ~1.5 orders of magnitude. The most proton-conductive compositions BaLa1.7Ba0.3In2O6.85 and BaLa1.7Sr0.15In2O6.925 have close conductivity values of 1.2 × 10−4 S/cm and 0.7 × 10−4 S/cm at 500 °C under wet air, respectively.

3. Materials and Methods

Two-layer perovskites BaLa2−xMxIn2O7−0.5x (M = Sr, Ba) were prepared by the solid-state method. The carbonates BaCO3 and SrCO3, and oxides In2O3 and In2O3 (99.99% purity, REACHIM, Moscow, Russia) were initially dried, then weighed (analytical balance (Sartorius AG, Göttingen, Germany)) and mixed in stoichiometric quantities. The agate mortar was used for milling the powders. The calcination was performed at 800, 900, 1000, 1100, 1200, and 1300 °C. The time of each temperature treatment was 24 h.
The X-ray analysis was performed using a Bruker Advance D8 Cu Kα diffractometer (Bruker, Billerica, MA, USA) with step of 0.01°, scanning rate of 0.5°/min.
The thermogravimetry (TG) and mass spectrometry (MS) analysis were performed using STA 409 PC Netzsch Analyser connected with QMS 403 C Aëolos mass spectrometer (Netzsch, Selb, Germany). The heating of initially hydrated samples was conducted in the temperature range of 40–1100 °C, with the rate of 10 °C/min under a flow of dry Ar.
The measurements of electrical conductivity were performed by impedance spectroscopy method (impedance spectrometer Z-1000P, (Electrochemical Instruments (Elins), Chernogolovka, Russia). The investigations were conducted from 1000 to 200 °C with 1°/min cooling rate under dry air or dry Ar conditions. The dry gas (air or Ar) was produced by circulating the gas through P2O5 (pH2O = 3.5·10−5 atm). The wet gas (air or Ar) 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 (Honeywell, Freeport, TX, USA). The dependencies of conductivities vs. partial oxygen pressures pO2 were obtained by using the electrochemical method for producing different pO2 with oxygen pump from Y-stabilized ZrO2 ceramic. The values of the resistance were recorded after 3–5 h of equilibrium.

4. Conclusions

It this paper, the members of the layered perovskites family with a two-layer Ruddlesden-Popper structure were obtained, and theirs transport properties were investigated. It is shown that doping of the lanthanum sublattice of BaLa2In2O7 by the ions Sr2+ and Ba2+ with bigger ionic radius leads to the increase in lattice parameters and unit cell volumes of doped compositions. The possibility for dissociative incorporation of water molecules into crystal lattice is proven. The water uptake is in the range of 0.15–0.22 mol water per formula unit of complex oxide for all undoped and doped compositions. In the dry air conditions, the nature of conductivity is mixed oxygen–hole, regardless of the dopant nature (Sr2+, Ba2+). Doping leads to the increase in the conductivity values by up to ~1.5 orders of magnitude. The most proton-conductive BaLa1.7Ba0.3In2O6.85 and BaLa1.7Sr0.15In2O6.925 samples are characterized by the conductivity values of 1.2·10−4 S/cm and 0.7·10−4 S/cm at 500 °C under wet air, respectively. The acceptor-doped layered perovskites based on BaLa2In2O7 are prospective proton-conducting materials, and further material science searches among this class of materials are relevant.

Author Contributions

Conceptualization, N.T. and I.A.; methodology, N.T. and I.A.; investigation, A.B., K.B., P.C. and E.A.; data curation, A.B., K.B., P.C. and E.A.; writing—original draft preparation, N.T. and I.A.; 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 received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ruddlesden, S.N.; Popper, P. New compounds of the K2NiF4 type. Acta Cryst. 1957, 10, 538–539. [Google Scholar] [CrossRef]
  2. Ruddlesden, S.N.; Popper, P. The compound Sr3Ti2O7 and its structure. Acta Cryst. 1958, 11, 54–55. [Google Scholar] [CrossRef] [Green Version]
  3. Aurivillius, B. Mixed bismuth oxides with layer lattices: I. Structure type of CaBi2B2O9. Arkiv Kem 1949, 1, 463–480. [Google Scholar]
  4. Dion, M.; Ganne, M.; Tournoux, M. Nouvelles familles de phases MIMII2Nb3O10 a feuillets «perovskites». Mater. Res. Bull. 1981, 16, 1429–1435. [Google Scholar] [CrossRef]
  5. Jacobson, A.J.; Lewandowski, J.T.; Johnson, J.W. Ion exchange of the layered perovskite KCa2Nb3O10 by protons. J. Less-Common Met. 1986, 116, 137–145. [Google Scholar] [CrossRef]
  6. Jacobson, A.J.; Johnson, J.W.; Lewandowski, J.T. Interlayer chemistry between thick transition-metal oxide layers: Synthesis and intercalation reactions of K[Ca2Nan−3NbnO3n+1]. Inorg. Chem. 1985, 24, 3727–3729. [Google Scholar] [CrossRef]
  7. Domen, K.; Yoshimura, J.; Sekine, T.; Tanaka, A.; Onishi, T. A novel series of photocatalysts with an ion-exchangeable layered structure of niobate. Catal Lett. 1990, 4, 339–343. [Google Scholar] [CrossRef]
  8. Machida, M.; Yabunaka, J.; Kijima, T. Efficient photocatalytic decomposition of water with the novel layered tantalate RbNdTa2O7. Chem. Commun. 1999, 1939–1940. [Google Scholar] [CrossRef]
  9. Machida, M.; Miyazaki, K.; Matsushima, S.; Arai, M. Photocatalytic properties of layered perovskite tantalates, MLnTa2O7 (M = Cs, Rb, Na, and H.; Ln = La, Pr, Nd, and Sm). J. Mater. Chem. 2003, 13, 1433–1437. [Google Scholar] [CrossRef]
  10. Rodionov, I.A.; Zvereva, I.A. Photocatalytic activity of layered perovskite-like oxides in practically valuable chemical reactions. Russ. Chem. Rev. 2016, 85, 248–279. [Google Scholar] [CrossRef]
  11. Krasheninnikova, O.V.; Syrov, E.V.; Smirnov, S.M.; Suleimanov, E.V.; Fukina, D.G.; Knyazev, A.V.; Titaev, D.N. Synthesis, crystal structure and photocatalytic activity of new Dion-Jacobson type titanoniobates. J. Solid State Chem. 2022, 315, 123445. [Google Scholar] [CrossRef]
  12. Phuruangrat, A.; Ekthammathat, N.; Dumrongrojthanath, P.; Thongtem, S.; Thongtem, T. Hydrothermal synthesis, structure, and optical properties of pure and silver-doped Bi2MoO6 nanoplates. Russ. J. Phys. Chem. 2015, 89, 2443–2448. [Google Scholar] [CrossRef]
  13. Chawla, H.; Chandra, A.; Ingole, P.P.; Garg, S. Recent advancements in enhancement of photocatalytic activity using bismuth-based metal oxides Bi2MO6 (M = W, Mo, Cr) for environmental remediation and clean energy production. Ind. Eng. Chem. Res. 2021, 95, 1–15. [Google Scholar] [CrossRef]
  14. Tasleem, S.; Tahir, M. Recent progress in structural development and band engineering of perovskites materials for photocatalytic solar hydrogen production: A review. Int. J. Hydrogen Energy 2020, 45, 19078–19111. [Google Scholar] [CrossRef]
  15. Huang, Y.; Liu, J.; Deng, Y.; Qian, Y.; Jia, X.; Ma, M.; Yang, C.; Liu, K.; Wang, Z.; Qu, S.; et al. The application of perovskite materials in solar water splitting. J. Semicond. 2020, 41, 011701. [Google Scholar] [CrossRef]
  16. Zhang, G.; Liu, G.; Wang, L.; Irvine, J.T.S. Inorganic perovskite photocatalysts for solar energy utilization. Chem. Soc. Rev. 2016, 45, 5951–5984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Zhang, P.; Zhang, J.; Gong, J. Tantalum-based semiconductors for solar water splitting. Chem. Soc. Rev. 2014, 43, 4395–4422. [Google Scholar] [CrossRef]
  18. Benedek, N.A.; Rondinelli, J.M.; Djani, H.; Ghosez, P.; Lightfoot, P. Understanding ferroelectricity in layered perovskites: New ideas and insights from theory and experiments. Dalton Trans. 2015, 44, 10543–10558. [Google Scholar] [CrossRef] [Green Version]
  19. Ferreira, W.C.; Rodrigues, G.L.C.; Araújo, B.S.; de Aguiar, F.A.A.; de Abreu Silva, A.N.A.; Fechine, P.B.A.; de Araujo Paschoal, C.W.; Ayala, A.P. Pressure-induced structural phase transitions in the multiferroic four-layer Aurivillius ceramic Bi5FeTi3O15. Ceram. Int. 2020, 46, 18056–180621. [Google Scholar] [CrossRef]
  20. Zulhadjri; Wendari, T.P.; Ikhram, M.; Putri, Y.E.; Septiani, U.; Imelda. Enhanced dielectric and ferroelectric responses in La3+/Ti4+ co-substituted SrBi2Ta2O9 Aurivillius phase. Ceram. Int. 2022, 48, 10328–103321. [Google Scholar] [CrossRef]
  21. Xu, Q.; Xie, S.; Wang, F.; Liu, J.; Shi, J.; Xing, J.; Chen, Q.; Zhu, J.; Wang, Q. Bismuth titanate based piezoceramics: Structural evolutions and electrical behaviors at different sintering temperatures. J. Alloys Compd. 2021, 882, 160637. [Google Scholar] [CrossRef]
  22. Chen, C.; Ning, H.; Lepadatu, S.; Cain, M.; Yan, H.; Reece, M.J. Ferroelectricity in Dion-Jacobson ABiNb2O7 (A = Rb, Cs) compounds. J. Mater. Chem. C 2015, 3, 19–22. [Google Scholar] [CrossRef] [Green Version]
  23. Kudo, A.; Kaneko, E. Photoluminescence of layered perovskite oxides with triple-octahedra slabs containing titanium and niobium. J. Mater. Sci. Lett. 1997, 16, 224–226. [Google Scholar] [CrossRef]
  24. Pavani, K.; Graça, M.P.F.; Kumar, J.S.; Neves, A.J. Photoluminescence varied by selective excitation in BiGdWO6:Eu3+ phosphor. Opt. Mater. 2017, 74, 120–127. [Google Scholar] [CrossRef]
  25. Mamidi, S.; Gundeboina, R.; Kurra, S.; Velchuri, R.; Muga, V. Aurivillius family of layered perovskites, BiREWO6 (RE = La, Pr, Gd, and Dy): Synthesis, characterization, and photocatalytic studies. Comptes Rendus Chim. 2018, 21, 547–552. [Google Scholar] [CrossRef]
  26. Zhou, G.; Jiang, X.; Zhao, J.; Molokeev, M.; Lin, Z.; Liu, Q.; Xia, Z. Two-dimensional-layered perovskite ALaTa2O7:Bi3+ (A = K and Na) phosphors with versatile structures and tunable photoluminescence. ACS Appl. Mater. Interfaces 2018, 10, 24648–24655. [Google Scholar] [CrossRef]
  27. Panda, D.P.; Singh, A.K.; Kundu, T.K.; Sundaresan, A. Visible-light excited polar Dion-Jacobson Rb(Bi1−xEux)2Ti2NbO10 perovskites: Photoluminescence properties and in vitro bioimaging. J. Mater. Chem. B 2022, 10, 935–944. [Google Scholar] [CrossRef]
  28. Yadav, D.; Nirala, G.; Kumar, U.; Upadhyay, S. Investigation on structural and optical properties of system Sr2Ce1−xNaxO4 (0.0 ≤ x ≤ 0.10). J. Mater. Sci. Mater. Electron. 2021, 32, 8064–8080. [Google Scholar] [CrossRef]
  29. Schaak, E.R.; Mallouk, T.E. KLnTiO4 (Ln = La, Nd, Sm, Eu, Gd, Dy): A new series of Ruddlesden-Popper phases synthesized by ion-exchange of HLnTiO4. J. Solid State Chem. 2001, 161, 225–232. [Google Scholar] [CrossRef] [Green Version]
  30. Nishimoto, S.; Matsuda, M.; Miyake, M. Novel protonated and hydrated n = 1 Ruddlesden-Popper phases, HxNa1−xLaTiO4−yH2O, formed by ion-exchange/intercalation reaction. J. Solid State Chem. 2005, 178, 811–818. [Google Scholar] [CrossRef]
  31. Nishimoto, S.; Matsuda, M.; Harjo, S.; Hoshikawa, A.; Kamiyama, T.; Ishigaki, T.; Miyake, M. Structural change in a series of protonated layered perovskite compounds, HLnTiO4 (Ln = La, Nd and Y). J. Solid State Chem. 2006, 179, 1892–1897. [Google Scholar] [CrossRef]
  32. Kurnosenko, S.A.; Silyukov, O.I.; Mazur, A.S.; Zvereva, I.A. Synthesis and thermal stability of new inorganic-organic perovskite-like hybrids based on layered titanates HLnTiO4 (Ln = La, Nd). Ceram. Int. 2020, 46, 5058–5068. [Google Scholar] [CrossRef]
  33. Huang, Y.; Xie, Y.; Fan, L.; Li, Y.; Wei, Y.; Lin, J.; Wu, J. Synthesis and photochemical properties of La-doped HCa2Nb3O10. Int. J. Hydrogen Energy 2008, 33, 6432–6438. [Google Scholar] [CrossRef]
  34. Zhou, C.; Shi, R.; Waterhouse, G.I.N.; Zhang, T. Recent advances in niobium-based semiconductors for solar hydrogen production. Coord. Chem. Rev. 2020, 419, 213399. [Google Scholar] [CrossRef]
  35. 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]
  36. 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]
  37. Ishihara, T.; Yan, Y.; Sakai, T.; Ida, S. Oxide ion conductivity in doped NdBaInO4. Solid State Ion. 2016, 288, 262–265. [Google Scholar] [CrossRef]
  38. Yang, X.; Liu, S.; Lu, F.; Xu, J.; Kuang, X. Acceptor doping and oxygen vacancy migration in layered perovskite NdBa1−nO4-based mixed conductors. J. Phys. Chem. C 2016, 120, 6416–6426. [Google Scholar] [CrossRef]
  39. 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]
  40. 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] [PubMed]
  41. Shiraiwa, M.; Kido, T.; Fujii, K.; Yashima, M. High-temperature proton conductors based on the (110) layered perovskite BaNdScO4. J. Mat. Chem. A 2021, 9, 8607. [Google Scholar] [CrossRef]
  42. Kato, S.; Ogasawara, M.; Sugai, M.; Nakata, S. Synthesis and oxide ion conductivity of new layered perovskite La1−xSr1+xInO4−d. Solid State Ion. 2002, 149, 53–57. [Google Scholar] [CrossRef]
  43. Troncoso, L.; Alonso, J.A.; Aguadero, A. Low activation energies for interstitial oxygen conduction in the layered perovskites La1+xSr1−xInO4+d. J. Mater. Chem. A 2015, 3, 17797–17803. [Google Scholar] [CrossRef]
  44. Troncoso, L.; Alonso, J.A.; Fernández-Díaz, M.T.; Aguadero, A. Introduction of interstitial oxygen atoms in the layered perovskite LaSrIn1−xBxO4+δ system (B=Zr, Ti). Solid State Ion. 2015, 282, 82–87. [Google Scholar] [CrossRef]
  45. Troncoso, L.; Mariño, C.; Arce, M.D.; Alonso, J.A. Dual oxygen defects in layered La1.2Sr0.8−xBaxInO4+d (x = 0.2, 0.3) oxide-ion conductors: A neutron diffraction study. Materials 2019, 12, 1624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Troncoso, L.; Arce, M.D.; Fernández-Díaz, M.T.; Mogni, L.V.; Alonso, J.A. Water insertion and combined interstitial-vacancy oxygen conduction in the layered perovskites La1.2Sr0.8−xBaxInO4+δ. New J. Chem. 2019, 43, 6087–6094. [Google Scholar] [CrossRef]
  47. Tarasova, N.; Animitsa, I.; Galisheva, A. Electrical properties of new protonic conductors Ba1+xLa1−xInO4−0.5x with Ruddlesden-Popper structure. J. Solid State Electrochem. 2020, 24, 1497–1508. [Google Scholar] [CrossRef]
  48. Tarasova, N.; Galisheva, A.; Animitsa, I. Improvement of oxygen-ionic and protonic conductivity of BaLaInO4 through Ti doping. Ionics 2020, 26, 5075–5088. [Google Scholar] [CrossRef]
  49. 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]
  50. 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]
  51. Tarasova, N.; Galisheva, A.; Animitsa, I.; Anokhina, I.; Gilev, A.; Cheremisina, P. Novel mid-temperature Y3+ → In3+ doped proton conductors based on the layered perovskite BaLaInO4. Ceram. Int. 2022, 48, 15677–15685. [Google Scholar] [CrossRef]
  52. Tarasova, N.; Animitsa, I.; Galisheva, A.; Korona, D.; Davletbaev, K. Novel proton-conducting layered perovskite based on BaLaInO4 with two different cations in B-sublattice: Synthesis, hydration, ionic (O2-, H+) conductivity. Int. J. Hydrogen Energy 2022, 47, 18972–18982. [Google Scholar] [CrossRef]
  53. 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]
  54. 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 Structur. Appl. Sci. 2022, 12, 4082. [Google Scholar] [CrossRef]
  55. 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. [Google Scholar] [CrossRef]
  56. Tarasova, N.; Galisheva, A.; Animitsa, I.; Korona, D.; Abakumova, E.; Medvedev, D. Novel mixed oxygen-electronic conductors based on BaLa2In2O7 with two-layer Ruddlesden-Popper structure. Ceram 2022, 48, 35376–35385. [Google Scholar] [CrossRef]
  57. 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 images of layered perovskites types of structures: Ruddlesden-Popper (An+1BnO3n+1), Dion–Jacobson (A’An−1BnO3n+1), and Aurivillius (An−1Bi2BnO3n+3) (from left to right) with n = 2.
Figure 1. The images of layered perovskites types of structures: Ruddlesden-Popper (An+1BnO3n+1), Dion–Jacobson (A’An−1BnO3n+1), and Aurivillius (An−1Bi2BnO3n+3) (from left to right) with n = 2.
Inorganics 10 00161 g001
Figure 2. The XRD data for the anhydrous BaLa1.8Ba0.2In2O6.9 (a) and hydrated BaLa1.8Ba0.2In2O6.9∙0.2H2O (b) compositions.
Figure 2. The XRD data for the anhydrous BaLa1.8Ba0.2In2O6.9 (a) and hydrated BaLa1.8Ba0.2In2O6.9∙0.2H2O (b) compositions.
Inorganics 10 00161 g002
Figure 3. The TG and MS(H2O) curves of BaLa1.8Ba0.2In2O6.9∙0.2H2O composition.
Figure 3. The TG and MS(H2O) curves of BaLa1.8Ba0.2In2O6.9∙0.2H2O composition.
Inorganics 10 00161 g003
Figure 4. The temperature dependencies of conductivity for the composition BaLa1.75Ba0.25In2O6.875 under dry (filled symbols) and wet (open symbols) conditions.
Figure 4. The temperature dependencies of conductivity for the composition BaLa1.75Ba0.25In2O6.875 under dry (filled symbols) and wet (open symbols) conditions.
Inorganics 10 00161 g004
Figure 5. The concentration dependencies of electrical conductivity at 500 °C under dry air for the solid solutions BaLa2−xMxIn2O7−0.5x (M = Sr, Ba).
Figure 5. The concentration dependencies of electrical conductivity at 500 °C under dry air for the solid solutions BaLa2−xMxIn2O7−0.5x (M = Sr, Ba).
Inorganics 10 00161 g005
Figure 6. The dependencies of the conductivity values vs. oxygen partial pressure for the compositions BaLa1.8Ba0.2In2O6.9 (a) and BaLa1.8Sr0.2In2O6.9 (b) under dry (filled symbols) and wet (open symbols) conditions and conductivity values from σ–103/T dependencies under wet air (filled black symbols) and wet Ar (open black symbols) conditions.
Figure 6. The dependencies of the conductivity values vs. oxygen partial pressure for the compositions BaLa1.8Ba0.2In2O6.9 (a) and BaLa1.8Sr0.2In2O6.9 (b) under dry (filled symbols) and wet (open symbols) conditions and conductivity values from σ–103/T dependencies under wet air (filled black symbols) and wet Ar (open black symbols) conditions.
Inorganics 10 00161 g006
Figure 7. The concentration dependencies of oxygen ion conductivity at 300, 500, and 700 °C for the solid solutions BaLa2−xBaxIn2O7−0.5x (a) and BaLa2−xSrxIn2O7−0.5x (b).
Figure 7. The concentration dependencies of oxygen ion conductivity at 300, 500, and 700 °C for the solid solutions BaLa2−xBaxIn2O7−0.5x (a) and BaLa2−xSrxIn2O7−0.5x (b).
Inorganics 10 00161 g007
Figure 8. The concentration dependencies of oxygen ions mobility at 700, 500, and 300 °C for the solid solutions BaLa2−xBaxIn2O7−0.5x (a) and BaLa2−xSrxIn2O7−0.5x (b).
Figure 8. The concentration dependencies of oxygen ions mobility at 700, 500, and 300 °C for the solid solutions BaLa2−xBaxIn2O7−0.5x (a) and BaLa2−xSrxIn2O7−0.5x (b).
Inorganics 10 00161 g008
Figure 9. The concentration dependencies of proton conductivity at 500 and 300 °C for the solid solutions BaLa2−xBaxIn2O7−0.5x (a) and BaLa2−xSrxIn2O7−0.5x (b).
Figure 9. The concentration dependencies of proton conductivity at 500 and 300 °C for the solid solutions BaLa2−xBaxIn2O7−0.5x (a) and BaLa2−xSrxIn2O7−0.5x (b).
Inorganics 10 00161 g009
Figure 10. The concentration dependencies of oxygen ions mobility at 300 °C for the solid solutions BaLa2−xBaxIn2O7−0.5x and BaLa2−xSrxIn2O7−0.5x.
Figure 10. The concentration dependencies of oxygen ions mobility at 300 °C for the solid solutions BaLa2−xBaxIn2O7−0.5x and BaLa2−xSrxIn2O7−0.5x.
Inorganics 10 00161 g010
Table 1. Lattice parameters, unit cell volume, and water uptake for BaLa2−xSrxIn2O7−0.5x.
Table 1. Lattice parameters, unit cell volume, and water uptake for BaLa2−xSrxIn2O7−0.5x.
Samplea, b (Å)c (Å)Vcell (Å3)Water Uptake (mol)
05.914(9)20.846(5)729.33650.17
0.055.915(2)20.869(0)730.19770.15
0.105.916(3)20.870(4)730.51830.18
0.155.916(4)20.871(3)730.57450.18
0.205.917(2)20.872(1)730.80010.19
Table 2. Lattice parameters, unit cell volume, and water uptake for BaLa2−xBaxIn2O7−0.5x.
Table 2. Lattice parameters, unit cell volume, and water uptake for BaLa2−xBaxIn2O7−0.5x.
Samplea, b (Å)c (Å)Vcell (Å3)Water Uptake (mol)
05.914(9)20.846(5)729.33650.17
0.055.915(1)20.859(0)729.82320.16
0.105.916(3)20.870(4)730.51830.17
0.155.920(4)20.899(3)732.54420.19
0.205.927(5)20.940(1)735.73570.20
0.255.941(5)20.949(1)739.53300.21
0.305.956(6)20.954(9)743.50250.22
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.; Bedarkova, A.; Animitsa, I.; Belova, K.; Abakumova, E.; Cheremisina, P.; Medvedev, D. Oxygen Ion and Proton Transport in Alkali-Earth Doped Layered Perovskites Based on BaLa2In2O7. Inorganics 2022, 10, 161. https://doi.org/10.3390/inorganics10100161

AMA Style

Tarasova N, Bedarkova A, Animitsa I, Belova K, Abakumova E, Cheremisina P, Medvedev D. Oxygen Ion and Proton Transport in Alkali-Earth Doped Layered Perovskites Based on BaLa2In2O7. Inorganics. 2022; 10(10):161. https://doi.org/10.3390/inorganics10100161

Chicago/Turabian Style

Tarasova, Nataliia, Anzhelika Bedarkova, Irina Animitsa, Ksenia Belova, Ekaterina Abakumova, Polina Cheremisina, and Dmitry Medvedev. 2022. "Oxygen Ion and Proton Transport in Alkali-Earth Doped Layered Perovskites Based on BaLa2In2O7" Inorganics 10, no. 10: 161. https://doi.org/10.3390/inorganics10100161

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