Next Article in Journal
Burapha-TH: A Multi-Purpose Character, Digit, and Syllable Handwriting Dataset
Next Article in Special Issue
Proton Conductivity of La2(Hf2xLax)O7x/2 “Stuffed” Pyrochlores
Previous Article in Journal
Zircon Hf-Isotopic Mapping Applied to the Metal Exploration of the Sanjiang Tethyan Orogenic Belt, Southwestern China
Previous Article in Special Issue
High-Temperature Behavior, Oxygen Transport Properties, and Electrochemical Performance of Cu-Substituted Nd1.6Ca0.4NiO4+δ Electrode Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Protonic Transport in Layered Perovskites BaLanInnO3n+1 (n = 1, 2) with Ruddlesden-Popper Structure

1
The Institute of High Temperature Electrochemistry of the Ural Branch of the Russian Academy of Sciences, 620066 Ekaterinburg, Russia
2
Institute of Natural Sciences and Mathematics, Ural Federal University, 620000 Yekaterinburg, Russia
*
Author to whom correspondence should be addressed.
Submission received: 17 January 2022 / Revised: 1 April 2022 / Accepted: 14 April 2022 / Published: 18 April 2022

Abstract

:
The work focused on the layered perovskite-related materials as the potential electrolytic components of such devices as proton conducting solid oxide fuel cells for the area of clean energy. The two-layered perovskite BaLa2In2O7 with the Ruddlesden–Popper structure was investigated as a protonic conductor for the first time. The role of increasing the amount of perovskite blocks in the layered structure on the ionic transport was investigated. It was shown that layered perovskites BaLanInnO3n+1 (n = 1, 2) demonstrate nearly pure protonic conductivity below 350 °C.

1. Introduction

The rising of the environmental challenges and the increasing in the need for safe and effective devices for energy conversion and storage mark the new decade [1,2,3,4]. This makes the task of creating high-efficiency solid oxide devices for electrochemical applications very important for now. The oxygen- and proton-conducting solid oxides may be components of various devices including fuel cells and electrolyzers [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21]. Most well-known electrolytes for medium-temperature solid oxide fuel cells (500–700 °C) are characterized by the perovskite or perovskite-related structure ABO3−δ [22]. As an example, one of the most investigated protonic conductors, barium zirconate [23], has the structure of ordinary perovskite with a cubic unit cell (Figure 1a). The varying of the cations A and B can create the materials with a layered perovskite-related structure where the perovskite layers alternate with rock-salt layers.
The general formula of layered perovskite-related structures can be written as AO(A′BO3)n, where n is the number of polyhedra layers stacking in the perovskite block. When n = 1, this formula can be represented as AA′BO4, and the compositions with this formula are characterized by the monolayered perovskite-related structure. The family of proton-conducting oxides with this type of the structure was discovered several years ago, and it is represented by the matrix compositions SrLaInO4 [24,25,26,27,28], BaLaInO4 [29,30,31,32,33,34,35,36,37], BaNdInO4 [38,39,40,41,42,43], BaNdScO4 [44]. The symmetry of these compositions is lower than cubic, and it is orthorhombic or monoclinic depending on the combination of cations in the A and B sublattices. As an example, the orthorhombic structure of the compound BaLaInO4 is represented in Figure 1b. It was shown that various types of doping (donor and acceptor) lead to the increase in the values of ionic conductivity [45]. Moreover, the compositions based on BaLaInO4 demonstrate almost pure protonic conductivity under wet air conditions at 450 °C [45].
The next member of the homologous series of the Ruddlesden–Popper type of structure with n = 2 can be writing in the general case as AA2′B2O7 and as an example BaLa2In2O7. Despite the possibility of protonic transport in the two-layered complex oxide, BaLa2In2O7 is not yet described; the structure of BaLanInnO3n+1 (n = 1, 2) is described in detail by Titov et al. [46] and Raveau et al. [47]. One of the main differences between BaLaInO4 and BaLa2In2O7 compositions is the ratio of perovskite blocks and rock-salt layers in the unit cell. This ratio is 1:1 for BaLaInO4 and 2:1 for BaLa2In2O7. It can be assumed that an increase in the ratio of perovskite blocks can lead to the increase in the ionic conductivity of solid oxide. Based on this, the goal of this work is the investigation of ionic transport (O2−, H+) in the two-layered compound BaLa2In2O7 in comparison with the mono-layered compound BaLaInO4. Proton transport is not previously reported for n = 2 RP phases in the Ba-La-In-O system.

2. Experimental

The phase BaLa2In2O7 was synthesized using the solid state method. The previously dried powder reagents BaCO3, La2O3, In2O3 were used as a starting materials. The milling of the reagents was made in an agate mortar. The calcination was performed in the temperature range from 800 to 1300 °C with the step of 100 °C and 24 h time treatments.
The XRD studies were performed using a Bruker Advance D8 (Rheinstetten, Germany) diffractometer with Cu Kα radiation over a range from 20° to 100° with a step of 0.01° and at a scanning rate of 0.5°/min. The preparation of the samples was made by their heating at 1100 °C for 5 h and following cooling in dry Ar (pH2O = 3.5 × 10−5 atm).
For the investigations of the electrical properties, the pressed cylindrical pellets (1300 °C, 24 h, dry air) were obtained. The pellets had a relative density ~92% (density of the sintered samples was determined by the Archimedes’ method). The ac conductivity measurements were performed by the Z-1000P (Elins, Russian Fedeartion) impedance spectrometer within the frequency range of 1–106 Hz. Electrical measurements were performed using Pt paste electrodes (annealing at 1000 °C for 2 h). The temperature dependencies of conductivities were obtained in the temperature range 200–1000 °C (step 10–20 °C, cooling rate of 1°/min; the measurements were taken until a constant resistance value was established). These 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 a 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 an oxygen pump (and sensor) from Y-stabilized ZrO2 ceramic. Each resistance value was recorded after 3–5 h of equilibrium.

3. Results and Discussions

The phase characterization of the obtained sample BaLa2In2O7 was made using XRD analysis. It was shown that all observed diffraction peaks can be related to the single phase belonging to the tetragonal symmetry (space group P42/mnm) (Figure 2). The lattice parameters (a = 5.914(9) Å, c = 20.846(5) Å) were in good agreement with those previously reported by Titov et al. and Raveau et al. (a = 5.915(2) Å, c = 20.86(1) Å [46]; a = 5.9141(3) Å, c = 20.8231(2) [47]). The structural parameters are presented in the Table 1.
The 3D visualization of the crystal structure of the obtained phase is presented in the Figure 1c using VESTA software [48]. As can be seen (Figure 1), for the mono-layered structure (Figure 1b), the ordering in the A and A′ cations does not happen, and Ba and La atoms occupy the same positions in the crystal lattice of BaLaInO4, and they have the same coordination number (c.n. = 9). Opposite, the ordered arrangement of A and A′ cations is observed for the two-layered complex oxide BaLa2In2O7, as it was shown by Raveau et al. [47] with using high resolution electron microscopy (Figure 1c). It was confirmed that barium atoms locate in the perovskite blocks, and the lanthanum atoms are placed in the rock-salt layers [47]. In the case of the two-layered structure, the cation with the largest ionic radii (barium in this case) has the c.n. = 12 and forms polyhedra [BaO12]; the cation with lower ionic radii (lanthanum) forms the polyhedra (LaO9) with c.n. = 9. However, it was shown that both perovskite blocks and rock-salt layers are distorted, and the real coordination numbers are ten and seven for barium and lanthanum, respectively [47].
The electrical conductivity was measured by the impedance spectroscopy method. The Nyquist-plots for the investigated composition BaLa2In2O7 obtained under dry air are presented in Figure 3a–c. The fitting of the spectra was made using ZView software, and the obtained results are presented in Table 2. According to the fitting of the spectra (red line) with using the equivalent circuit presented in Figure 3d, three different electrochemical processes can be defined. For the calculation of electrical conductivity, the bulk resistance values R1 were used and are discussed below. It can be noted that, due to the small depression of the semicircles, the constant phase element (CPE) was used during the analysis of Nyquist plots.
The temperature dependencies of bulk conductivities obtained under dry air and dry Ar are presented in Figure 4. As can be seen, the conductivity values for the compound BaLa2In2O7 are higher than for BaLaInO4 [29] by around 0.6–0.7 order of magnitude. This confirms the previously stated assumption that the increase in the ratio of perovskite blocks in the layered structures can lead to the increase in the electrical conductivity values. However, for the correct comparison, the total conductivity values must be divided into ionic and electronic contributions.
To establish the nature of electrical transport in BaLa2In2O7, the conductivity measurements vs. pO2 were performed (Figure 5). As can be seen (Figure 5a), in the electrolytic area (pO2 = 10−16–10−5 atm), the independence of conductivity values from pO2 was observed. Thus, the oxygen-ionic conductivity dominates in this region. For layered structures, oxygen-ionic transport is usually 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.
The possibility of the formation of interstitial oxygen is provided by the existence of a rock-salt block and the possibility of increasing the coordination number of the cation that is located there. Accordingly, the formation of oxygen vacancies occurs in the perovskite block; the cation, octahedrally surrounded by oxygen, lowers the coordination number. That is, the presence of oxygen-deficient perovskite blocks plays an important role in the formation of oxygen-ion transport. Therefore, for such structures, it is possible to expect a significant oxygen-ionic conductivity without doping. Moreover, the measurements of the conductivity versus pO2 showed that the phases of BaLa2In2O7 are stable over a wide range of pO2.
The conductivity dependencies obtained in the dry oxidizing conditions (pO2 = 10−5–0.21 atm) have a positive slope, which indicates the mixed ionic-electronic (hole) nature of conductivity:
1 2 O 2 O i + 2 h
V o + 1 2 O 2 O O x + h
where O i is the oxygen atom in the interstitial position; V O   is the oxygen vacancy; h is the hole. Obviously, both of these processes can be realized in the structure. However, it is impossible to say which process prevails. For the layered composition based on BaNdInO4, the calculations of formation energies of point defects were made [41]. The energy of processes described by Equation (3) is about 1.2 eV. However, the data for Equation (2) are missing.
It should be noted that the conductivity values obtained in the dry Ar (blue symbols in the Figure 5) are the same as the conductivity values from the electrolytic area (black symbols in the Figure 5a). Therefore, in an argon atmosphere, it is relatively easy to measure the temperature dependence of oxygen-ionic conductivity in a wide temperature range.
It was shown earlier, for the compositions based on BaLaInO4, the conductivity values obtained under dry Ar corresponded to the oxygen-ionic conductivity values [45]. Based on this, 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 both BaLa2In2O7 and BaLaInO4 compositions did not change under temperature variation at pO2 = 0.21 atm and were around 20% for each composition, and both compositions demonstrate mixed ionic-electronic (hole) conductivity under dry air, as it was shown earlier for the composition BaLaInO4 [45].
Returning to Figure 4a, it can be seen that the values of oxygen-ionic conductivity for BaLa2In2O7 were higher than for BaLaInO4. It should the noted that the increasing in the oxygen-ionic conductivity values correlates well with the decreasing in the values of energy activation of oxygen transport (from 0.87 eV for BaLaInO4 to 0.81 eV for BaLaInO4). Because the increase in the oxygen-ionic conductivity correlates with the increase in the number of perovskite layers in the structure, we can suppose that the transport of oxygen ions includes the oxygen positions in the In-containing polyhedra.
The comparison of the dependencies of conductivity vs. oxygen partial pressure under dry and wet conditions is presented in Figure 5b. At the pO2 range from 10−4 to 10−16 atm, the conductivity increases significantly in a wet atmosphere as a result of the appearance of a proton contribution due to the dissolution of H2O in BaLa2In2O7.
Water molecules can be dissolved in the oxygen-deficient perovskite layers leading to the formation of proton carriers as follows:
V o + H 2 O + O o × 2 OH o
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.
At the same time, for layered structures containing rock-salt blocks, the second mechanism of the dissociative dissolution of water molecules is possible:
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 rock-salt block. This process is accompanied by the increase in the coordination number of atoms in the rock-salt block.
The increase in the conductivity in a wet atmosphere compared to a dry atmosphere is reduced with increasing temperature. In a wet atmosphere below 450 °C, the conductivity does not depend on pO2, which reflects the dominant ionic (proton) nature of the conductivity.
Differences in the conductivities in dry and wet atmospheres at high pO2 > 10−4 atm are not so significant compared to the plateau region (Figure 5b). This is due to the presence of holes, the concentration of which decreases in a wet atmosphere:
h + 1 2 H 2 O   + O i 1 4 O 2 + ( OH ) i
where h is the hole; O i is the oxygen atom in the interstitial position; ( OH ) i is the oxygen-hydrogen group.
In accordance with the explanations presented above and the obtained data on conductivity-pO2, the measurements of conductivity in a wide temperature range can be carried out by comparing the atmosphere of air and argon at different levels of humidity. This comparison is shown in Figure 4.
As can be seen (Figure 4b), the values of ionic conductivity (wet Ar) are near the same as the total conductivity values (wet air) below ~450 °C. In other words, the nature of conductivity changes from mixed ionic-hole (dry air) to predominantly ionic (wet air, T < 450 °C).
The difference between conductivity values obtained under atmospheres with different water partial pressures is well visible for the Ar atmosphere, and these temperature dependencies are presented in Figure 4c. The protonic transport numbers tp were calculated according to the equation:
t p = σ w e t   A r σ d r y   A r σ w e t
which was used earlier by Y. Zhou et al. [43] for the layered perovskites based on BaNdInO4. It was established that for both compositions BaLa2In2O7 and BaLaInO4, the protonic transport numbers increase with a decreasing temperature and reach ~90–95% in the temperature range 300–350 °C.
Therefore, the possibility of protonic transport in the two-layered complex oxide BaLa2In2O7 with the Ruddlesden–Popper structure was shown for the first time. It was proved that an increase in n in the formula BaLanInnO3n+1 leads to the increase in the ionic conductivity. The composition BaLa2In2O7 demonstrates nearly pure protonic conductivity under wet air below 350 °C. This allows for considering this complex oxide as a promising matrix composition for obtaining novel high-conductive protonic electrolytes based on it.

4. Conclusions

In this work, the complex oxide BaLa2In2O7 was synthesized, and the electrical conductivity was measured for the first time. It was shown that change in the ratio between perovskite blocks and rock-salt layers in the layered perovskite-related complex oxides BaLanInnO3n+1 (n = 1, 2) influences the ionic conductivity values. The increasing n in the formula leads to the increase in the ionic conductivity due to the increase in the share of perovskite blocks. It was proved that the complex oxide BaLa2In2O7 demonstrates nearly pure protonic conductivity under wet air below 350 °C, and it can be considered as the promising compound for obtaining novel high-conductive protonic electrolytes based on it.

Author Contributions

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

Funding

This research received no external funding.

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. Corvalan, C.; Prats, E.V.; Sena, A.; Varangu, L.; Vinci, S. Towards Climate Resilient and Environmentally Sustainable Health Care Facilities. Int. J. Environ. Res. Public Health 2020, 17, 8849. [Google Scholar] [CrossRef]
  2. Watts, N.; Amann, M.; Arnell, N.; Montgomery, H.; Costello, A. The 2020 report of The Lancet Countdown on health and climate change: Responding to converging crises. Lancet 2021, 397, 129–170. [Google Scholar] [CrossRef]
  3. Afroze, S.; Reza, M.S.; Cheok, Q.; Taweekun, J.; Azad, A.K. Solid oxide fuel cell (SOFC); A new approach of energy generation during the pandemic COVID-19. Int. J. Integr. Eng. 2020, 12, 245–256. [Google Scholar] [CrossRef]
  4. Afroze, S.; Reza, M.S.; Cheok, Q.; Islam, S.N.; Abdalla, A.M.; Taweekun, J.; Azad, A.K.; Khalilpoor, N.; Issakhov, A. Advanced Applications of Fuel Cells during the COVID-19 Pandemic. Int. J. Chem. Eng. 2021, 2021, 5539048. [Google Scholar] [CrossRef]
  5. Sun, C.; Alonso, J.A.; Bian, J. Recent Advances in Perovskite-Type Oxides for Energy Conversion and Storage Applications. Adv. Energy Mater. 2020, 11, 2000459. [Google Scholar] [CrossRef]
  6. Shi, H.; Su, C.; Ran, R.; Cao, J.; Shao, Z. Electrolyte materials for intermediate-temperature solid oxide fuel cells. Prog. Nat. Sci. Mater. Int. 2020, 30, 764–774. [Google Scholar] [CrossRef]
  7. Yang, B.; Guo, Z.; Wang, J.; Wang, J.; Zhu, T.; Shu, H.; Qiu, G.; Chen, J.; Zhang, J. Solid oxide fuel cell systems fault diagnosis: Critical summarization, classification, and perspectives. J. Energy Storage 2021, 34, 102153. [Google Scholar] [CrossRef]
  8. Peng, J.; Huang, J.; Wu, X.-L.; Xu, Y.-W.; Chen, H.; Li, X. Solid oxide fuel cell (SOFC) performance evaluation, fault diagnosis and health control: A review. J. Power Sources 2021, 5051, 230058. [Google Scholar] [CrossRef]
  9. 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]
  10. Singh, M.; Zappa, D.; Comini, E. Solid oxide fuel cell: Decade of progress, future perspectives and challenges. Int. J. Hydrogen Energy 2021, 46, 27643–276745. [Google Scholar] [CrossRef]
  11. Shen, M.; Ai, F.; Ma, H.; Xu, H.; Zhang, Y. Progress and prospects of reversible solid oxide fuel cell materials. iScience 2021, 24, 103464. [Google Scholar] [CrossRef]
  12. 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, 5024–5037. [Google Scholar] [CrossRef]
  13. 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]
  14. 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]
  15. Medvedev, D. Trends in research and development of protonic ceramic electrolysis cells. Int. J. Hydrogen Energy 2019, 44, 26711–26740. [Google Scholar] [CrossRef]
  16. Shim, J.H. Ceramics breakthrough. Nat. Energy 2018, 3, 168–169. [Google Scholar] [CrossRef]
  17. 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]
  18. 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]
  19. Zvonareva, I.; Fu, X.-Z.; Medvedev, D.; Shao, Z. Electrochemistry and energy conversion features of protonic ceramic cells with mixed ionic-electronic electrolytes. Energy Environ. Sci. 2022, 15, 439–465. [Google Scholar] [CrossRef]
  20. 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]
  21. 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. 2021, 46, 2212–2240. [Google Scholar] [CrossRef]
  22. Irvine, J.; Rupp, J.L.; Liu, G.; Xu, X.; Haile, S.; Qian, X.; Snyder, A.; Freer, R.; Ekren, D.; Skinner, S. Roadmap on inorganic perovskites for energy applications. J. Phys. Energy 2021, 3, 031502. [Google Scholar] [CrossRef]
  23. Hossain, M.K.; Chanda, R.; El-Denglawey, A.; Emrose, T.; Rahman, M.T.; Biswas, M.C.; Hashizume, K. Recent progress in barium zirconate proton conductors for electrochemical hydrogen device applications: A review. Ceram. Int. 2021, 47, 23725–23748. [Google Scholar] [CrossRef]
  24. 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]
  25. 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]
  26. 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]
  27. 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]
  28. 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]
  29. Tarasova, N.; Animitsa, I.; Galisheva, A.; Korona, D. Incorporation and Conduction of Protons in Ca, Sr, Ba-Doped BaLaInO4 with Ruddlesden-Popper Structure. Materials 2019, 12, 1668. [Google Scholar] [CrossRef] [Green Version]
  30. Tarasova, N.; Animitsa, I.; Galisheva, A.; Pryakhina, V. Protonic transport in the new phases BaLaIn0.9M0.1O4.05 (M=Ti, Zr) with Ruddlesden-Popper structure. Solid State Sci. 2020, 101, 106121. [Google Scholar] [CrossRef]
  31. 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]
  32. 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]
  33. 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]
  34. 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]
  35. 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]
  36. 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 Comm. 2021, 323, 14093. [Google Scholar] [CrossRef]
  37. 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]
  38. 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]
  39. 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]
  40. Ishihara, T.; Yan, Y.; Sakai, T.; Ida, S. Oxide ion conductivity in doped NdBaInO4. Solid State Ion. 2016, 288, 262–265. [Google Scholar] [CrossRef]
  41. 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]
  42. 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]
  43. 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]
  44. 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]
  45. 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]
  46. 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]
  47. 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]
  48. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
Figure 1. Structure of BaZrO3 (ABO3) (a), BaLaInO4 (AA′BO4) (b) and BaLa2In2O7 (AA2′B2O7) (c), where red spheres represent the oxygen atoms; green spheres represent the atoms of A-sublattice; and yellow spheres represent the atoms of A′-sublattice.
Figure 1. Structure of BaZrO3 (ABO3) (a), BaLaInO4 (AA′BO4) (b) and BaLa2In2O7 (AA2′B2O7) (c), where red spheres represent the oxygen atoms; green spheres represent the atoms of A-sublattice; and yellow spheres represent the atoms of A′-sublattice.
Applsci 12 04082 g001
Figure 2. The XRD-data for the composition BaLa2In2O7.
Figure 2. The XRD-data for the composition BaLa2In2O7.
Applsci 12 04082 g002
Figure 3. The Nyquist-plots obtained at the different temperatures under dry air for the composition BaLa2In2O7: (a) 700 °C, (b) 600 °C, (c) 500 °C, and (d) the equivalent circuit of fitting (red line).
Figure 3. The Nyquist-plots obtained at the different temperatures under dry air for the composition BaLa2In2O7: (a) 700 °C, (b) 600 °C, (c) 500 °C, and (d) the equivalent circuit of fitting (red line).
Applsci 12 04082 g003
Figure 4. The temperature dependencies of conductivity of BaLa2In2O7 (blue symbols) and BaLaInO4 (black symbols) [29] obtained dry Ar (open symbols) and dry air (closed symbols) (a), wet Ar (open symbols) and wet air (closed symbols) (b), wet Ar (open symbols), and dry Ar (closed symbols) (c).
Figure 4. The temperature dependencies of conductivity of BaLa2In2O7 (blue symbols) and BaLaInO4 (black symbols) [29] obtained dry Ar (open symbols) and dry air (closed symbols) (a), wet Ar (open symbols) and wet air (closed symbols) (b), wet Ar (open symbols), and dry Ar (closed symbols) (c).
Applsci 12 04082 g004
Figure 5. The dependencies of the conductivity values vs. pO2 for the sample BaLa2In2O7 at dry conditions (a) and dry (closed symbols) and wet (open symbols) conditions (b) (black symbols) and conductivity values from σ–103/T dependencies at dry Ar (blue symbols).
Figure 5. The dependencies of the conductivity values vs. pO2 for the sample BaLa2In2O7 at dry conditions (a) and dry (closed symbols) and wet (open symbols) conditions (b) (black symbols) and conductivity values from σ–103/T dependencies at dry Ar (blue symbols).
Applsci 12 04082 g005
Table 1. Refined structural parameters from XRD-data.
Table 1. Refined structural parameters from XRD-data.
AtomSitexyz
Ba4f0.2483(1)0.2483(1)0
La8j0.2687(4)0.2687(4)0.1851(3)
In8j0.2590(0)0.2590(0)0.4002(2)
O(1)4g0.806(1)0.193(3)0
O(2)8j0.183(4)0.183(4)0.291(1)
O(3)8h00.50.095(2)
O(4)4e000.126(1)
O(5)4e000.382(1)
Rp = 4.31, Rwp = 3.48, χ2 = 1.73.
Table 2. Results of Nyquist-plots fitting, where CPE is the constant phase element (F), and R is the resistance (kΩ∙cm).
Table 2. Results of Nyquist-plots fitting, where CPE is the constant phase element (F), and R is the resistance (kΩ∙cm).
ElementValue (700 °C)Value (600 °C)Value (500 °C)
CPE12.1 × 10−122.4 × 10−123.2 × 10−12
R11336158
CPE24.1 × 10−103.7 × 10−102.1 × 10−10
R22833
CPE36.0 × 10−75.6 × 10−77.0 × 10−7
R33739
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.; 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. https://0-doi-org.brum.beds.ac.uk/10.3390/app12084082

AMA Style

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. Applied Sciences. 2022; 12(8):4082. https://0-doi-org.brum.beds.ac.uk/10.3390/app12084082

Chicago/Turabian Style

Tarasova, Nataliia, Anzhelika Galisheva, Irina Animitsa, Daniil Korona, Hala Kreimesh, and Irina Fedorova. 2022. "Protonic Transport in Layered Perovskites BaLanInnO3n+1 (n = 1, 2) with Ruddlesden-Popper Structure" Applied Sciences 12, no. 8: 4082. https://0-doi-org.brum.beds.ac.uk/10.3390/app12084082

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