Next Article in Journal
Influence of a Nano-Hydrophobic Admixture on Concrete Durability and Steel Corrosion
Previous Article in Journal
Comparative Manufacturing of Hybrid Composites with Waste Graphite Fillers for UAVs
Previous Article in Special Issue
Modernized Synthesis Technique of Pr2NiO4+δ-Based Complex Oxides Using Low-Temperature Salt Melts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Advanced Proton-Conducting Ceramics Based on Layered Perovskite BaLaInO4 for Energy Conversion Technologies and Devices

by
Nataliia Tarasova
1,2,* and
Anzhelika Bedarkova
1,2
1
Laboratory of Electrochemical Devices Based on Solid Oxide Proton Electrolytes, The Institute of High Temperature Electrochemistry of the Ural Branch of the Russian Academy of Sciences, 620660 Yekaterinburg, Russia
2
Institute of Hydrogen Energy, Ural Federal University, 620002 Yekaterinburg, Russia
*
Author to whom correspondence should be addressed.
Materials 2022, 15(19), 6841; https://doi.org/10.3390/ma15196841
Submission received: 8 September 2022 / Revised: 23 September 2022 / Accepted: 27 September 2022 / Published: 1 October 2022
(This article belongs to the Special Issue Development and Application of Solid Oxide Electrolytes)

Abstract

:
Production of high efficiency renewable energy source for sustainable global development is an important challenge for humans. Hydrogen energy systems are one of the key elements for the development of sustainable energy future. These systems are eco-friendly and include devices such as protonic ceramic fuel cells, which require advanced proton-conducting materials. In this study, we focused on new ceramics with significantly improved target properties for hydrogen energy purposes. Neodymium-doped phase based on layered perovskite BaLaInO4 was obtained for the first time. The ability for water intercalation and proton transport was proved. It was shown that the composition BaLa0.9Nd0.1InO4 is the predominant proton conductor below 400 °C under wet air. Moreover, isovalent doping of layered perovskites AA′BO4 is the promising method for improving transport properties and obtaining novel advanced proton-conducting ceramic materials.

1. Introduction

The goal of sustainable global development is an important challenge of human society for many decades [1,2,3,4,5,6,7,8]. The creation and production of high efficiency renewable energy source is a major opportunity for this achievement [9,10,11,12,13,14,15]. Hydrogen can be considered as one of the key elements for the development of sustainable energy future [16,17,18,19]. The eco-friendly hydrogen-production and hydrogen-operation systems include devices such as protonic ceramic fuel cells and protonic ceramic electrolysis cells, which require advanced proton-conducting materials [20,21,22,23,24,25,26,27,28,29,30,31,32]. Such type of solid-state conductors was discovered in early 1980s by Iwahara et al. and these phases were derivatives from strontium cerate [33,34,35]. These complex oxides are characterized by the perovskite structure, and many of recent developed protonic conductors have related structure, for example structure of double perovskites [36] or hexagonal perovskites [37,38].
At the same time, the development of hydrogen energy devices requires not only the creation of novel materials with improved properties but the solution of the problem of materials comparability with each other in electrochemical device also. Today, one of the promising cathode materials is layered perovskite based on Ln2NiO4+δ [39,40,41,42]. Likely enough, the similarity of crystal structure can help to resolve the problem of comparability of electrode and electrolyte components. By this way, the materials search of advanced proton conductors characterized by the layered perovskite structures is on the focus.
Layered perovskites with general formula AA′BO4 where A is Ba or Sr, A′ is La or Nd, and B is In or Sc were described as ionic (oxygen-ionic and protonic) conductors in the past few years [43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62]. It was proved that doping in the cationic sublattices is the successful way for improving protonic conductivity up to ~1.5 orders. At the same time, most of the works are related to the methods of heterovalent doping. Doping by the alkali-earth cations such as Ca2+, Sr2+, Ba2+ into La3+ (acceptor doping) [49] sublattice and Ti4+, Zr4+, Nb4+ into In3+ sublattice [50] (donor doping) was investigated. However, isovalent doping shows even more meaningful results. Earlier, we reported about the possibility of isovalent doping of B-sublattice of layered perovskite BaLaInO4 by the ions with different radii (Y, Sc) [54,55,56]. In this paper, we focused on the modification of BaLaInO4 structure by the isovalent doping in the A (lanthanum) sublattice (Figure 1). Neodymium was chosen as a trivalent metal with a close radius to lanthanum. The Nd-doped composition BaLa0.9Nd0.1InO4 was studied as the proton conductor for the first time. In addition, the comparative analysis with previously obtained data was carried out. The effect of dopant nature on the crystal lattice parameters and on the mobility of ions was revealed.

2. Materials and Methods

Layered perovskite BaLa0.9Nd0.1InO4 was prepared by the solid-state method. The carbonates BaCO3, CaCO3 and oxides Nd2O3, In2O3 were used.
The X-ray analysis was made using a Bruker Advance D8 diffractometer (Bruker, Billerica, MA, USA). Morphology of the powder sample was defined by Phenom ProX (ThermoFisher, Waltham, MA, USA) Desktop scanning electron microscope (SEM) integrated with energy-dispersive X-ray diffraction (EDS) detector.
The thermogravimetry (TG) analysis was made using STA 409 PC Netzsch Analyzer (Netzsch, Selb, Germany). The initially hydrated samples were used for the investigations.
The electrical conductivity was measured using impedance spectrometer Z-1000P (Elins, RF). The samples were cooled from 1000 to 200 °C under dry (pH2O = 3.5 × 10−5 atm) or wet (pH2O = 2 × 10−2 atm) conditions. Dry gas was obtained by circulating through P2O5. The wet gas was obtained by bubbling through saturated solution of KBr (pH2O = 2 × 10−2 atm).

3. Results

3.1. X-ray, SEM, and EDS Characterization

XRD-analysis of powder sample BaLa0.9Nd0.1InO4 confirmed the single-phase nature of obtained composition. Both previously obtained and investigated scandium-doped BaLaIn0.9Sc0.1O4 and yttrium-doped BaLaIn0.9Y0.1O4 compositions, and the neodymium-doped sample crystallize in the Pbca space group (Figure 2a) and isostructural to the matrix composition BaLaInO4. The results of Rietveld refinement of BaLa0.9Nd0.1InO4 composition are presented in the Figure 3a and Table 1 The changes in lattice parameters of these compositions during doping compared with BaLaInO4 (a = 12.932(3) Å, b = 5.906(0) Å, c = 5.894(2) Å) are presented in the Figure 2b and Table 2. As can be seen, despite of “plus” or “minus” difference in the ionic radii of the metals ( r In 3 + = 0.80 Å, r Sc 3 +   = 0.745 Å, r Y 3 + = 0.90 Å, r La 3 + = 1.216 Å, r Nd 3 + = 1.163 Å [63]), doping led to an increase in the a lattice parameter (interlayer space) for all compositions. At first sight, increase in the lattice parameter during doping by the ions with smaller ionic radii (Nd3+→La3+ and Sc3+→In3+) can be considered as contradiction. However, this is seeming contradiction, which can be explained on closer examination.
It is obvious that doping led to the formation substitution defects MLa and MIn. From the quasi-chemical point of view, they are neutral and can be written as Nd La × , Sc In × , and Y In × . On the contrary, from the crystallochemical point of view, these defects can be written as Nd La δ , Sc In δ + , and Y In δ + because of difference in the electronegativity of elements ( χ In = 1.78, χ Sc = 1.36, χ Y = 1.22, χ La = 1.10, χ Nd = 1.14 [64]). In other words, the redistribution of the electron density and the change in the effective charges on the atoms take place, which leads to changes in the energy and length of metal-oxygen bonds. Accordingly, the change in the lattice parameters during doping may not have a direct correlation with the size of the dopant. We can conclude that observing one-way trend of increase in a lattice parameter is due to the appearance of additional repulsion effects of different nature ions in one sublattice. It should be noted that in contrast to the perovskite 3D structure where the octahedrons are connected by all six vertices, the layered perovskite structure contains the octahedra layers bonded only by axial oxygens and non-bonded by apical oxygens. Thus, this structure is more flexible and more easily able to change the crystal-chemical distances in comparison with the classic perovskite structure.
Verification of chemical composition of obtained sample BaLa0.9Nd0.1InO4 was performed using SEM coupled with energy-dispersive diffraction analysis. Good agreement between theoretical and obtained values is observed (Table 3). Sample consists of irregularly rounded grains ~3–5 µm, forming agglomerates up to 15 µm (Figure 3).

3.2. TG-Measurements

For the materials with classic perovskite structure ABO3–δ, the possibility of water uptake depends on the amount of oxygen vacancies in the structure and is determined by the value of δ in the general case. However, for the layered perovskite AA′BO4, this process is provided by the intercalation of OH-groups into the space between perovskite layers (inset in the Figure 4):
H 2 O + O O x ( OH ) O + ( OH   ) i
Moreover, the amount of water uptake for heterovalent-doped samples based on BaLaInO4 does not depend on the concentration of oxygen vacancies, but it is determined by the unit cell volume of the composition [62]. The water uptake area for the acceptor- and donor-doped compositions based on BaLaInO4 is presented in the Figure 4 in blue color. As can be seen, the water uptake for the isovalent-doped compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4, and BaLaIn0.9Y0.1O4 is well correlated with the value of the unit cell volume also. Thus, we can conclude that the doping mechanism does not affect the possibility of water uptake for compositions based on BaLaInO4. The only factor determining the amount of water absorption is the unit cell volume of the layered perovskite.
The TG-curve for the BaLa0.9Nd0.1InO4 composition in comparison with curves for previously reported undoped and scandium- and yttrium-doped compositions are presented in the Figure 5. All samples are dehydrated in several steps. The appearance of several steps on the TG-curve confirms the presence in the structure of non-equivalent protons characterized by different thermal stabilities. The dehydration temperature is higher when OH-groups are more strongly bonded to the crystal lattice and are less involved in hydrogen bonds. Low-temperature protons (200–350 °C) are removed first and corresponded to the strongly bonded hydroxyl groups [55]. High-temperature protons (350–700 °C) are removed later and corresponded to the weakly bonded or relatively isolated hydroxyl groups. As can be seen (inset in the Figure 5), the share of low-temperature protons decreases with the increase in the unit cell volume of doped compositions.

3.3. Electrical Properties

The impedance spectroscopy method was used for the investigation of electrical properties. The Nyquist plots for BaLa0.9Nd0.1InO4 composition are presented in the Figure 6. Fitting of experimental data (red line) was made. The applied equivalent circuit is presented in the Figure 6, where R1 is the bulk resistance, R2 is the grain boundaries resistance. The bulk resistance values R1 were used for the calculation of electrical conductivity (Table 4).
Figure 7 represents the temperature dependencies of conductivities for BaLa0.9Nd0.1InO4 composition. The conductivity obtained at dry Ar (filled red symbols) is lower than value obtained under dry air (filled blue symbols) in whole investigated temperature range. As it was shown earlier for the undoped and doped materials based on BaLaInO4, the conductivity in dry Ar (~10–5 atm) is conductivity of oxygen-ionic conductivity [51,52,54]. Thus, we can say that the composition BaLa0.9Nd0.1InO4 is characterized by the mixed oxygen ionic-hole conductivity in dry air. The oxygen-ionic transport numbers can be calculated as:
t O 2 = σ O 2 σ tot = σ Ar σ air
and they are about 40% in the whole temperature range. At the same time, the t O 2 for the undoped composition is about 20% [49] and they are also independent of the temperature. Thus, doping leads to the increase in total conductivity and share of oxygen-ionic conductivity.
The relation of conductivity at wet air (open blue symbols) and at wet Ar (open red symbols) is different in comparison with dry condition. At high temperatures (higher than 450 °C), the conductivity values in wet air are higher compared to values in wet Ar. At low temperatures (below 450 °C), the conductivity in wet air is comparable with the conductivity in wet Ar. Because proton concentration becomes significant at low temperatures, the increase in conductivity in wet Ar compared to dry Ar is due to the formation of protonic species which leads to a decrease in the contribution of the whole conductivity:
h + 1 2 H 2 O   + O i 1 4 O 2 + ( OH ) i
The protonic conductivity was calculated as the difference between conductivity in wet and dry Ar. The proton transport numbers tp were calculated according to the equation:
t p = σ wet   Ar σ dry   Ar σ wet
They were about 95% below 400 °C for the composition BaLa0.9Nd0.1InO4. The same protonic conductivity transport numbers were obtained earlier for the BaLaIn0.9Sc0.1O4 [54] and BaLaIn0.9Y0.1O4 [55] compositions also.
The protons mobility values were calculated according to the formula:
μ H + = σ H + e c H +
The temperature dependencies of proton conductivity and mobility for the BaLa0.9Nd0.1InO4 composition in comparison with curves for previously reported undoped, scandium-doped, and yttrium-doped compositions are presented in the Figure 8a,b correspondingly. The isovalent doping leads not only to an increase in the protonic conductivity but also proton mobility. Figure 9 demonstrates protonic conductivity and mobility vs. a lattice parameter for the undoped BaLaInO4 and isovalent-doped compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4, and BaLaIn0.9Y0.1O4 at 350 °C. Both conductivity and mobility of protons increase with an increasing in a lattice parameter for all isovalent-doped compositions. In the other words, the increase in protonic conductivity is provided not only by the changes in the concentration of protons, but also by an increase in their mobility. It is obvious that this increase in mobility is due to the expansion of the interlayer space (a lattice parameter) which facilitates the transport of protons.
Figure 10 represents the same dependencies of protonic conductivity and mobility vs. a lattice parameter for isovalent-doped compositions in comparison with heterovalent-doped compositions with the same (0.1 mol) dopant content (colored areas in the Figure 10). As can be seen, the isovalent doping allows to increase more significantly the conductivity and mobility of protons in comparison with heterovalent doping at the same dopant concentration. In the other words, the mobility of protons in the heterovalent-doped compositions is lower in comparison with isovalent-doped compositions at the same value of a lattice parameter. Thus, there is an additional factor affecting the mobility of protons in the layered perovskites. As it was shown for the heterovalent-doped compositions [62], the mobility of protons decreased at “high” (more ~ 0.1 mol) dopant concentrations due to the formation of proton-aggregating clusters:
M A + ( OH ) o M A · ( OH ) o ×
M B + ( OH ) i M B · ( OH ) i ×
It is obvious that the process of cluster formation occurs at lower concentrations of the dopant also. Opposite to the heterovalent doping, the isovalent doping can lead to the indirect formation of partially charged defects ( Nd La δ , Sc In δ + , Y In δ + ) only. Accordingly, the share of proton-containing clusters is significantly lower or even absent. This suggests that isovalent doping of layered perovskites AA′BO4 is a more promising method for improving transport properties compared with heterovalent doping. This method can be applied for obtaining novel advanced proton-conducting ceramics which can be used as electrolytic material in different energy conversion devices.

4. Conclusions

In this paper, neodymium-doped phase based on layered perovskite BaLaInO4 was obtained for the first time. The ability for water intercalation and proton transport was proved. It was shown that the composition BaLa0.9Nd0.1InO4 is the predominant proton conductor in the wet air below 400 °C. The comparative analysis of transport properties of isovalent-doped and heterovalent-doped protonic conductors based on BaLaInO4 was also carried out. First, all types of doping leads to an increase in the protonic conductivity values. Second, the increase in the protonic conductivity is provided not only by an increase in the concentration of protons, but also by an increase in their mobility. The increase in mobility is due to the expansion of the interlayer space which facilitates the transport of protons. Third, the isovalent doping allows to increase more significantly the conductivity and mobility of protons in comparison with heterovalent doping at the same dopant concentration. Isovalent doping of layered perovskites AA′BO4 is the promising method for improving transport properties and obtaining novel advanced proton-conducting ceramic materials.

Author Contributions

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

Funding

This research was supported by the Russian Science Foundation (grant no 22-79-10003).

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. Holdren, J.P.; Ehrlich, P.R. Human population and the global environment. Am. Sci. 1974, 62, 282–292. [Google Scholar]
  2. Norgaard, R.B. Sustainable development: A co-evolutionary view. Futures 1988, 20, 606–620. [Google Scholar] [CrossRef]
  3. Winter, C.-J.; Nitsch, J. Hydrogen energy-a “sustainable development” towards a world energy supply system for future decades. Int. J. Hydrog. Energy 1989, 14, 785–796. [Google Scholar] [CrossRef]
  4. Kats, G.H. Slowing global warming and sustaining development: The promise of energy efficiency. Energy Policy 1990, 18, 25–33. [Google Scholar] [CrossRef]
  5. Kumar, A.; Reddy, N. Energy for sustainable development. Int. J. Glob. Energy 1990, 2, 143–147. [Google Scholar] [CrossRef]
  6. Dincer, I. Renewable energy and sustainable development: A crucial review. Renew. Sustain. Energy Rev. 2000, 4, 157–175. [Google Scholar] [CrossRef]
  7. Stambouli, A.B.; Traversa, E. Solid oxide fuel cells (SOFCs): A review of an environmentally clean and efficient source of energy. Renew. Sustain. Energy Rev. 2002, 6, 433–455. [Google Scholar] [CrossRef]
  8. Panwar, N.L.; Kaushik, S.C.; Kothari, S. Role of renewable energy sources in environmental protection: A review. Renew. Sustain. Energy Rev. 2011, 15, 1513–1524. [Google Scholar] [CrossRef]
  9. Chu, S.; Majumdar, A. Opportunities and challenges for a sustainable energy future. Nature 2012, 48, 294–303. [Google Scholar] [CrossRef]
  10. Dincer, I.; Rosen, M.A. Sustainability aspects of hydrogen and fuel cell systems. Int. J. Sustain. Energy Dev. 2011, 15, 137–146. [Google Scholar] [CrossRef]
  11. Branco, H.; Castro, R.; Lopes, A.S. Battery energy storage systems as a way to integrate renewable energy in small isolated power systems. Int. J. Sustain. Energy Dev. 2018, 43, 90–99. [Google Scholar] [CrossRef]
  12. Malerba, D. Poverty-energy-emissions pathways: Recent trends and future sustainable development goals. Int. J. Sustain. Energy Dev. 2019, 49, 109–124. [Google Scholar] [CrossRef]
  13. Buonomano, A.; Barone, G.; Forzano, C. Advanced energy technologies, methods, and policies to support the sustainable development of energy, water and environment systems. Energy Rep. 2022, 8, 4844–4853. [Google Scholar] [CrossRef]
  14. Olabi, A.G.; Abdelkareem, M.A. Renewable energy and climate change. Renew. Sustain. Energy Rev. 2022, 158, 112111. [Google Scholar] [CrossRef]
  15. Østergaard, P.A.; Duic, N.; Noorollahi, Y.; Mikulcic, H.; Kalogirou, S. Sustainable development using renewable energy technology. Renew. Energy 2020, 146, 2430–2437. [Google Scholar] [CrossRef]
  16. International Energy Agency. The Future of Hydrogen: Seizing Today’s Opportunities; OECD: Paris, France, 2019. [Google Scholar] [CrossRef]
  17. Abdalla, A.M.; Hossain, S.; Nisfindy, O.B.; Azad, A.T.; Dawood, M.; Azad, A.K. Hydrogen production, storage, transportation and key challenges with applications: A review. Energy Convers. Manag. 2018, 165, 602–627. [Google Scholar] [CrossRef]
  18. Dawood, F.; Anda, M.; Shafiullah, G.M. Hydrogen production for energy: An overview. Int. J. Hydrog. Energy 2020, 45, 3847–3869. [Google Scholar] [CrossRef]
  19. Arsad, A.Z.; Hannan, M.A.; Al-Shetwi, A.Q.; Mansur, M.; Mittaqi, K.M.; Dong, Z.Y.; Blaabjerg, F. Hydrogen energy storage integrated hybrid renewable energy systems: A review analysis for future research directions. Int. J. Hydrog. Energy 2022, 47, 17285–17312. [Google Scholar] [CrossRef]
  20. Hossain, S.; Abdalla, A.M.; Jamain, S.N.B.; Zaini, J.H.; Azad, A.K. A review on proton conducting electrolytes for clean energy and intermediate temperature-solid oxide fuel cells. Renew. Sustain. Energy Rev. 2017, 79, 750–764. [Google Scholar] [CrossRef]
  21. Kim, J.; Sengodan, S.; Kim, S.; Kwon, O.; Bu, 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]
  22. Zhang, W.; Hu, Y.H. Progress in proton-conducting oxides as electrolytes for low-temperature solid oxide fuel cells: From materials to devices. Energy Sci. Eng. 2021, 9, 984–1011. [Google Scholar] [CrossRef]
  23. 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]
  24. Medvedev, D. Trends in research and development of protonic ceramic electrolysis cells. Int. J. Hydrog. Energy 2019, 44, 26711–26740. [Google Scholar] [CrossRef]
  25. 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]
  26. 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]
  27. Shim, J.H. Ceramics breakthrough. Nat. Energy 2018, 3, 168–169. [Google Scholar] [CrossRef]
  28. 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]
  29. Irvine, J.; Rupp, J.L.M.; Liu, G.; Xu, X.; Haile, S.; Qian, X.; Snyder, A.; Freer, R.; Ekren, D.; Skinner, S.; et al. Roadmap on inorganic perovskites for energy applications. J. Phys. Energy 2021, 3, 031502. [Google Scholar] [CrossRef]
  30. Chiara, A.; Giannici, F.; Pipitone, C.; Longo, A.; Aliotta, C.; Gambino, M.; Martorana, A. Solid-Solid Interfaces in Protonic Ceramic Devices: A Critical Review. ACS Appl. Mater. Interfaces 2020, 12, 55537–55553. [Google Scholar] [CrossRef]
  31. Cao, J.; Ji, Y.; Shao, Z. New Insights into the Proton-Conducting Solid Oxide Fuel Cells. J. Chin. Ceram. Soc. 2021, 49, 83–92. [Google Scholar] [CrossRef]
  32. Bello, I.T.; Zhai, S.; Zhao, S.; Li, Z.; Yu, N.; Ni, M. Scientometric review of proton-conducting solid oxide fuel cells. Int. J. Hydrog. Energy 2021, 46, 37406–37428. [Google Scholar] [CrossRef]
  33. Iwahara, H.; Esaka, T.; Uchida, H.; Maeda, N. Proton conduction in sintered oxides and its application to steam electrolysis for hydrogen production. Solid State Ion. 1981, 3–4, 359–363. [Google Scholar] [CrossRef]
  34. Iwahara, H.; Uchida, H.; Maeda, N. High temperature fuel and steam electrolysis cells using proton conductive solid electrolytes. J. Power Sources 1982, 7, 293–301. [Google Scholar] [CrossRef]
  35. Iwahara, H.; Uchida, H.; Tanaka, S. High temperature type proton conductors based on SrCeO3 and its application to solid electrolyte fuel cells. Solid State Ion. 1983, 9–10, 1021–1025. [Google Scholar] [CrossRef]
  36. Tarasova, N.; Colomban, P.; Animitsa, I. The short-range structure and hydration process of fluorine-substituted double perovskites based on barium-calcium niobate Ba2CaNbO5.5. J. Phys. Chem. Solids 2018, 118, 32–39. [Google Scholar] [CrossRef]
  37. Fop, S.; McCombie, K.S.; Wildman, E.J.; Skakle, J.M.S.; Irvine, J.T.S.; Connor, P.A.; Savaniu, C.; Ritter, C.; McLaughlin, A.C. High oxide ion and proton conductivity in a disordered hexagonal perovskite. Nat. Mater. 2020, 19, 752–757. [Google Scholar] [CrossRef]
  38. Yashima, M.; Tsujiguchi, T.; Sakuda, Y.; Yasui, Y.; Zhou, Y.; Fujii, K.; Torii, S.; Kamiyama, T.; Skinner, S.J. High oxide-ion conductivity through the interstitial oxygen site in Ba7Nb4MoO20-based hexagonal perovskite related oxides. Nat. Comm. 2021, 12, 556. [Google Scholar] [CrossRef]
  39. Tarutin, A.; Lyagaeva, J.; Medvedev, D.; Bi, L.; Yaremchenko, A. Recent advances in layered Ln2NiO4+δ nickelates: Fundamentals and prospects of their applications in protonic ceramic fuel and electrolysis cells. J. Mater. Chem. A 2021, 9, 154–195. [Google Scholar] [CrossRef]
  40. Tarutin, A.; Gorshkov, M.Y.; Bainov, A.; Vdovin, G.; Vylkov, A.; Lyagaeva, J.; Medvedev, D. Barium-doped nickelates Nd2–xBaxNiO4+δ as promising electrode materials for protonic ceramic electrochemical cells. Ceram. Int. 2020, 46, 24355–24364. [Google Scholar] [CrossRef]
  41. Tarutin, A.; Lyagaeva, J.; Farlenkov, A.; Plaksin, S.; Vdovin, G.; Demin, A.; Medvedev, D. A Reversible Protonic Ceramic Cell with Symmetrically Designed Pr2NiO4+δ-Based Electrodes: Fabrication and Electrochemical Features. Materials 2019, 12, 118. [Google Scholar] [CrossRef]
  42. Tarutin, A.P.; Lyagaeva, J.G.; Farlenkov, A.S.; Vylkov, A.I.; Medvedev, D.A. Cu-substituted La2NiO4+δ as oxygen electrodes for protonic ceramic electrochemical cells. Ceram. Int. 2019, 45, 16105–16112. [Google Scholar] [CrossRef]
  43. 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]
  44. 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]
  45. Ishihara, T.; Yan, Y.; Sakai, T.; Ida, S. Oxide ion conductivity in doped NdBaInO4. Solid State Ionics 2016, 288, 262–265. [Google Scholar] [CrossRef]
  46. 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]
  47. Fujii, K.; Yashima, M. Discovery and development of BaNdInO4 -A brief review. J. Ceram. Soc. Jpn. 2018, 126, 852–859. [Google Scholar] [CrossRef] [Green Version]
  48. 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]
  49. 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]
  50. 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]
  51. 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]
  52. 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]
  53. Tarasova, N.; Galisheva, A.; Animitsa, I. Ba2+/Ti4+- co-doped layered perovskite BaLaInO4: The structure and ionic (O2−, H+) conductivity. Int. J. Hydrog. Energy 2021, 46, 16868–16877. [Google Scholar] [CrossRef]
  54. 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]
  55. 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]
  56. 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 Hydrog. Energy 2022, 47, 18972–18982. [Google Scholar] [CrossRef]
  57. 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]
  58. 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]
  59. 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]
  60. 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]
  61. 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+d. New J. Chem. 2019, 43, 6087–6094. [Google Scholar] [CrossRef]
  62. 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]
  63. 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]
  64. Allred, A.L. Electronegativity values from thermochemical data. J. Inorg. Nucl. Chem. 1961, 17, 215–221. [Google Scholar] [CrossRef]
Figure 1. The isovalent substitution representation of BaLaInO4.
Figure 1. The isovalent substitution representation of BaLaInO4.
Materials 15 06841 g001
Figure 2. The XRD-patterns of isovalent-doped compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4, BaLaIn0.9Y0.1O4 (a) and lattice parameters changes of these compositions compared with BaLaInO4 (b).
Figure 2. The XRD-patterns of isovalent-doped compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4, BaLaIn0.9Y0.1O4 (a) and lattice parameters changes of these compositions compared with BaLaInO4 (b).
Materials 15 06841 g002
Figure 3. The refinement of XRD-pattern (a) and SEM-images of the compositions BaLa0.9Nd0.1InO4 (b).
Figure 3. The refinement of XRD-pattern (a) and SEM-images of the compositions BaLa0.9Nd0.1InO4 (b).
Materials 15 06841 g003
Figure 4. The dependency of water uptake vs. unit cell volume for the compositions BaLa0.9Nd0.1InO4 (Nd), BaLaIn0.9Sc0.1O4 (Sc) [54], BaLaIn0.9Y0.1O4 (Y) [55], BaLaInO4 (In) [49]. The blue area defines the water uptake area for the acceptor- and donor-doped compositions based on BaLaInO4 with 0.1 mol dopant content [62]. Schematic illustration of water insertion into BaLaInO4 structure is also shown.
Figure 4. The dependency of water uptake vs. unit cell volume for the compositions BaLa0.9Nd0.1InO4 (Nd), BaLaIn0.9Sc0.1O4 (Sc) [54], BaLaIn0.9Y0.1O4 (Y) [55], BaLaInO4 (In) [49]. The blue area defines the water uptake area for the acceptor- and donor-doped compositions based on BaLaInO4 with 0.1 mol dopant content [62]. Schematic illustration of water insertion into BaLaInO4 structure is also shown.
Materials 15 06841 g004
Figure 5. The TG-data for the compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4 [54], BaLaIn0.9 Y0.1O4 [55], BaLaInO4 [49]. The shares of protons with different thermal stabilities for the doped compositions are also shown.
Figure 5. The TG-data for the compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4 [54], BaLaIn0.9 Y0.1O4 [55], BaLaInO4 [49]. The shares of protons with different thermal stabilities for the doped compositions are also shown.
Materials 15 06841 g005
Figure 6. The Nyquist plots obtained at 350, 400, and 450 °C under wet air for the composition BaLa0.9Nd0.1InO4.
Figure 6. The Nyquist plots obtained at 350, 400, and 450 °C under wet air for the composition BaLa0.9Nd0.1InO4.
Materials 15 06841 g006
Figure 7. The conductivity dependencies form temperature for BaLa0.9Nd0.1InO4 composition.
Figure 7. The conductivity dependencies form temperature for BaLa0.9Nd0.1InO4 composition.
Materials 15 06841 g007
Figure 8. Dependencies of protonic conductivity (a) and mobility (b) from temperature for the compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4 [54], BaLaIn0.9Y0.1O4 [55], BaLaInO4 [49].
Figure 8. Dependencies of protonic conductivity (a) and mobility (b) from temperature for the compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4 [54], BaLaIn0.9Y0.1O4 [55], BaLaInO4 [49].
Materials 15 06841 g008
Figure 9. Dependencies of protonic conductivity (a) and mobility (b) from a lattice parameter for the compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4 [54], BaLaIn0.9Y0.1O4 [55], BaLaInO4 [49] at 350 °C.
Figure 9. Dependencies of protonic conductivity (a) and mobility (b) from a lattice parameter for the compositions BaLa0.9Nd0.1InO4, BaLaIn0.9Sc0.1O4 [54], BaLaIn0.9Y0.1O4 [55], BaLaInO4 [49] at 350 °C.
Materials 15 06841 g009
Figure 10. Dependencies of protonic conductivity (a) and mobility (b) from a lattice parameter for the compositions BaLa0.9Nd0.1InO4 (Nd), BaLaIn0.9Sc0.1O4 (Sc) [54], BaLaIn0.9Y0.1O4 (Y) [55], BaLaInO4 (In) [49] at 350 °C. The orange and yellow areas are defined as the protonic conductivity and mobility areas correspondingly for the acceptor- and donor-doped compositions based on BaLaInO4 with 0.1 mol dopant content [62].
Figure 10. Dependencies of protonic conductivity (a) and mobility (b) from a lattice parameter for the compositions BaLa0.9Nd0.1InO4 (Nd), BaLaIn0.9Sc0.1O4 (Sc) [54], BaLaIn0.9Y0.1O4 (Y) [55], BaLaInO4 (In) [49] at 350 °C. The orange and yellow areas are defined as the protonic conductivity and mobility areas correspondingly for the acceptor- and donor-doped compositions based on BaLaInO4 with 0.1 mol dopant content [62].
Materials 15 06841 g010
Table 1. Refined structural parameters from XRD-data.
Table 1. Refined structural parameters from XRD-data.
AtomSitexyz
Ba8c0.1439(2)−0.0028(3)0.0031(1)
La8c0.1439(2)−0.0028(3)0.0031(1)
In4b0.500
O(1)8c0.0193(1)0.203(3)0.203(3)
O(2)8c0.332(4)0.0193(10.0095(4)
Rp = 4.22, Rwp = 3.31, χ2 = 1.78.
Table 2. The parameters of unit cell for the isovalent-doped compositions based on BaLaInO4.
Table 2. The parameters of unit cell for the isovalent-doped compositions based on BaLaInO4.
CompositionBaLa0.9Nd0.1InO4BaLaIn0.9Sc0.1O4 [54]BaLaIn0.9Y0.1O4 [55]
a, Å12.948(5)12.951(9)12.969(1)
b, Å5.907(2)5.895(1)5.883(2)
c, Å5.903(5)5.883(2)5.911(6)
V, Å3451.55(7)449.19(8)455.24(7)
Table 3. The average element ratios determined by EDS analysis for the sample BaLa0.9Nd0.1InO4 (theoretical values are in brackets).
Table 3. The average element ratios determined by EDS analysis for the sample BaLa0.9Nd0.1InO4 (theoretical values are in brackets).
ElementBaLaNdIn
Content, atomic %33.6 (33.3)29.7 (30.0)3.2 (3.3)33.5 (33.4)
Table 4. The Nyquist plots fitting results.
Table 4. The Nyquist plots fitting results.
ElementValue (450 °C)Value (400 °C)Value (350 °C)
CPE1, F2.3 × 10−122.1 × 10−123.1 × 10−12
R1, kΩ92746
CPE2, F4.2 × 10−103.8 × 10−102.2 × 10−10
R2, kΩ134
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. Advanced Proton-Conducting Ceramics Based on Layered Perovskite BaLaInO4 for Energy Conversion Technologies and Devices. Materials 2022, 15, 6841. https://doi.org/10.3390/ma15196841

AMA Style

Tarasova N, Bedarkova A. Advanced Proton-Conducting Ceramics Based on Layered Perovskite BaLaInO4 for Energy Conversion Technologies and Devices. Materials. 2022; 15(19):6841. https://doi.org/10.3390/ma15196841

Chicago/Turabian Style

Tarasova, Nataliia, and Anzhelika Bedarkova. 2022. "Advanced Proton-Conducting Ceramics Based on Layered Perovskite BaLaInO4 for Energy Conversion Technologies and Devices" Materials 15, no. 19: 6841. https://doi.org/10.3390/ma15196841

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