Next Article in Journal / Special Issue
The Effect of the Formation of Superelastic NiTi Phase on Static and Dynamic Corrosion Performance of Ni-P Coating
Previous Article in Journal / Special Issue
Dynamics and Elastic Properties of Glassy Metastable States
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Protonated Forms of Layered Perovskite-Like Titanate NaNdTiO4: Neutron and X-ray Diffraction Structural Analysis

Department of Chemical Thermodynamics and Kinetics, Institute of Chemistry, Saint Petersburg State University, 198504 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Submission received: 20 April 2021 / Revised: 22 June 2021 / Accepted: 23 June 2021 / Published: 1 July 2021
(This article belongs to the Special Issue Feature Papers of Solids 2021)

Abstract

:
Structures of partially and completely protonated Ruddlesden–Popper phases, H0.7Na0.3NdTiO4·0.3H2O and HNdTiO4, have been established by means of neutron and X-ray diffraction analysis and compared among themselves as well as with that of the initial titanate NaNdTiO4. It was shown that while interlayer sodium cations in the partially protonated form are coordinated by nine oxygen atoms, including one related to intercalated water, in the fully protonated compound the ninth oxygen proves to be an axial anion belonging to the opposite slab of titanium-oxygen octahedra. Moreover, the partially protonated titanate was found to significantly differ from the other two in the octahedron distortion pattern. It is characterized by a weakly pronounced elongation of the octahedra towards the Nd-containing interlayer space making Ti4+ cations practically equidistant from both axial oxygen atoms, which is accompanied by a low-frequency shift of the bands relating to the asymmetric stretching mode of axial Ti–O bonds observed in the Raman spectra.

1. Introduction

Over past decades, great attention has been paid to ion-exchangeable layered perovskite-like oxides as promising ionic conductors [1,2,3,4,5], catalysts [6], photocatalysts [7,8,9], multipurpose precursors for the creation of hybrid inorganic–organic materials [10,11,12,13,14,15,16,17,18], nanoelectronic [19], ferroelectric [20], piezoelectric [21], luminescent [22] and photovoltaic devices [23], chemical sensors [24] as well as fuel cells [25].
Ion-exchangeable layered perovskites are crystalline solids possessing a block-type structure, which consists of negatively charged intergrowth slabs with a thickness of n corner-shared perovskite octahedra regularly alternating with interlayer spaces populated by alkali cations. Their valuable physical-chemical properties are associated with the unique perovskite structure as well as the pronounced activity of the interlayer space towards ion exchange and intercalation reactions [26,27]. In accordance with features of structure and composition, ion-exchangeable layered perovskites may be divided into two classes: Dion–Jacobson [28,29] and Ruddlesden–Popper phases [30,31]. The Dion–Jacobson phases follow a general formula A’[An−1BnO3n+1], where A’ is an alkali metal, A is alkaline earth or transition metal and B is Nb, Ta, Ti, Zr or related element. The relative arrangement of adjacent perovskite slabs in these compounds depends on the interlayer cation size. In particular, the Dion–Jacobson perovskites with comparatively small cations (A’ = Li, Na) are usually characterized by a staggered conformation while in the case of larger ones (A’ = K, Rb, Cs) an eclipsed conformation is more probable. Due to relatively low interlayer charge density, the Dion–Jacobson phases demonstrate high reactivity in a wide range of low-temperature transformations [32,33,34,35]. The structure of the Ruddlesden–Popper phases, described by a formula A’2[An−1BnO3n+1], represents alternation of perovskite slabs with the layers adopting the crystal arrangement of rock salt. These substances are characterized by higher interlayer charge density and their adjacent perovskite slabs are usually stacked in a staggered conformation. The reactivity of the Ruddlesden–Popper perovskites is known to be lower in comparison with that of the Dion–Jacobson ones, although it appears to be enough for ion exchange and intercalation of small molecules via the soft chemistry approaches [36,37,38,39,40,41].
A deep understanding of the structural features of layered perovskite-like oxides is of great importance for the prediction of their physical-chemical properties and targeted research-based creation of materials with specified characteristics. For instance, ionic conductivity strongly depends on the equilibrium crystallographic positions of interlayer cations associated in turn with a degree of the perovskite octahedra distortion [42]. As well, the interlayer space is considered a separate reaction zone in heterogeneous photocatalysis providing superior photocatalytic activity of many layered perovskites in comparison with that of non-layered materials due to reversible reactants’ intercalation [43]. In this regard, its structural features may significantly influence the supply of reactants and the withdrawal of products defining the reaction kinetics and, therefore, the photocatalyst performance. Moreover, the hydration degree and, apparently, crystallographic positions of intercalated water molecules affect the reactivity of layered perovskites with organic compounds [44]. Knowing these peculiarities greatly facilitates the creation of hybrid inorganic-organic materials for many promising applications.
Layered perovskite-like titanates ALnTiO4 (A = alkali cation or proton, Ln = La or lanthanoid cation) are Ruddlesden–Popper phases whose unique structure is characterized by the complete ordering of the aforementioned cations between two nonequivalent interlayer spaces interbedded by perovskite slabs with a thickness of n = 1 titanium-oxygen octahedron experiencing pronounced vertical distortion [45,46]. These compounds attract the attention of researchers as promising proton conductors [47,48,49,50,51], catalysts [52], photocatalysts [53,54,55,56] as well as luminescent [57] and magnetically susceptible materials [58,59,60].
As shown in our previous reports [61,62], when placed in aqueous solutions, the alkaline titanates ALnTiO4 undergo ion exchange and hydration giving protonated hydrated forms HxA1−xLnTiO4·yH2O whose specific composition depends on the pH of the medium and a particular interlayer cation. Since many practically valuable processes, such as photocatalytic reactions, are conducted in aqueous media, a thorough investigation of water-stable forms of layered perovskites deserves special attention. In view of the foregoing, the present paper focuses on a structural study of two protonated forms of the layered perovskite-like titanate NaNdTiO4, namely the partially protonated form H0.7Na0.3NdTiO4·0.3H2O, existing in water excess, and the completely protonated one HNdTiO4, being stable in acid media.

2. Materials and Methods

2.1. Synthesis

The alkaline layered perovskite-like titanate NaNdTiO4 was synthesized according to the standard ceramic technique using preliminarily calcined TiO2, Nd2O3 and Na2CO3 as reactants. The oxides were taken in stoichiometric amounts, sodium carbonate—with a 30% excess to compensate for the loss during calcination. All the compounds were mixed and thoroughly ground in an agate mortar under a layer of n-hexane for 30 min per each gram. The mixture obtained was dried and pelletized into ~2 g tablets at 50 bar using an Omec PI 88.00 hydraulic press. The tablets were placed into corundum crucibles with lids, kept at 850 °C for 12 h in a Nabertherm L-011K2RN muffle furnace and, after cooling down, ground in an agate mortar.
To obtain the partially protonated hydrated form H0.7Na0.3NdTiO4·0.3H2O, 1 g of NaNdTiO4 was treated with 1000 mL of water at room temperature for 1 d. In the case of the completely protonated form HNdTiO4 preparation, 1 g of NaNdTiO4 was processed with 200 mL of a 0.1 M hydrochloric acid solution under the same conditions. Then the solid phases were separated from solutions via centrifuging and dried in a desiccator over CaO for 2 d.

2.2. Instrumentation

Powder X-ray diffraction (XRD) patterns of the samples were recorded on a Thermo Scientific ARL X’TRA diffractometer (CuKα radiation, angle range = 5–120°, accumulation time 10 h). The lattice parameters calculation and Rietveld refinement were performed based on all the reflections observed using GSAS software. The initial structural parameters were adopted from the paper [63]. Neutron diffraction patterns were obtained on an IR-8 research neutron reactor (neutron wavelength λ = 1.7526 Å, angle range = 10–130°, accumulation time 50 h). The samples were investigated in aluminum containers with a diameter of 3 mm and a wall thickness of 0.03 mm whose reduced diffraction peaks were subsequently subtracted from the patterns. The structure calculations were conducted using GSAS software. The initial parameters for the refinement were taken from the reports [63,64]. Thermogravimetric (TG) analysis was carried out on a Netzsch TG 209 F1 Libra thermobalance in an argon atmosphere. The temperature program included heating each sample from room temperature to 700 °C at a rate of 10 °C/min. The protonation degree and amount of intercalated water were determined according to the previously reported method [65]. Raman scattering spectra were collected on a Bruker Senterra spectrometer in the Raman shift range of 100–4000 cm−1 using the incident laser 488 nm (power 20 mW, accumulation time 10 s). The high-frequency region in the figures was presented on an enlarged scale to make bands of water more visible relative to the high-intensity ones of the perovskite matrix.

3. Results and Discussion

Initial alkaline titanate NaNdTiO4 and its protonated forms were identified using XRD analysis (Figure 1). Tetragonal lattice parameters of NaNdTiO4 (a = b = 3.75 Å c = 12.82 Å) and HNdTiO4 (a = b = 3.70 Å c = 12.09 Å) were found to be consistent with the literature values [46]. In the course of the partial protonation, the interlayer distance d is seen to increase (this is reflected in the shift of low-angle 00l reflections on the diffractogram to the region of lower values), which indicates probable water intercalation accompanying the ion exchange. Otherwise, the d value should have reduced since protons are much smaller than sodium cations. Moreover, the partially protonated form is amenable to indexing in another space group (I4/mmm instead of P4/nmm) and is characterized by doubled c parameter, which is not observed in the case of NaNdTiO4 and HNdTiO4. The space group change and the c parameter doubling appear to be caused by a probable transformation of the adjacent perovskite slabs conformation from staggered to eclipsed. The completely protonated titanate HNdTiO4, apparently, possesses the slabs arrangement similar to that of NaNdTiO4 and does not contain a noticeable amount of interlayer water.
TG curves of the protonated samples are presented in Figure 2. In both cases, two main sections on the curves can be distinguished. The first of them corresponding to the temperatures up to 200 °C refers to deintercalation of the interlayer water molecules resulting in the formation of an anhydrous partially or fully protonated titanate. At the second section (250–650 °C), the latter experiences topochemical dehydration consisting in the binding of the interlayer oxygen atoms and protons into water molecules, their liberation and closing vertices of titanium-oxygen octahedra of adjacent perovskite slabs. In this way, final thermolysis products follow a gross formula Na1−xNdTiO4−x/2, where x is the amount of protons in the initial compound. Processing of the curves allowed us to establish quantitative compositions signed in Figure 2. The partially protonated titanate contains a perceptible amount of intercalated water released during heating at low temperatures. It corresponds to the formula H0.7Na0.3NdTiO4·0.3H2O (specified H0.73Na0.27NdTiO4·0.3H2O). The completely protonated HNdTiO4 was found to exist in a practically anhydrous state.
The structural model of the partially protonated titanate H0.7Na0.3NdTiO4·0.3H2O was refined by the Rietveld method based on the XRD data (Figure 3) [62]. The XRD patterns collected made it possible to determine with high accuracy the positions of all the atoms constituting the compound, besides hydrogen (Table 1). All the parameters calculated fall in a physically reasonable range (although the Na occupancy factor and position are hardly reliable) and are consistent with the calculation data for the related La-containing titanate. Thus, the refinement of the H0.7Na0.3NdTiO4·0.3H2O structure can be considered successful.
An attempt to refine the neutron diffraction pattern was made for the fully protonated form (Figure 4). In accordance with the results obtained (Table 2), the determination of the HNdTiO4 structure proved to be successful. All the parameters being refined take valid values and the difference between experimental and calculated profiles lies within the limits admissible for structural calculations.
Data on the comparison of HNdTiO4, the structure of which was refined here, and its deuterated counterpart DNdTiO4, the structure of which was refined in one of the earlier reports [64], are shown in Table 3 and Figure 5. Generally, the structural models of both compounds can be considered practically identical. The deuterated sample demonstrates a slightly greater lattice parameter in comparison with those of the non-deuterated one. In the latter protons are located closer to the axial oxygen that binds them than deuterium cations (0.919 Å vs. 0.932 Å). At the same time, the titanium-oxygen octahedron distortion is not proportional to the lattice parameter increasing. In particular, the bond Ti–O3 proves to be 0.01 Å longer in the deuterated titanate.
As noted earlier, the complete substitution of sodium cations by protons leads to a noticeable decrease of the distance between vertices of the ion-exchangeable layer octahedra. However, a space group change or a relative shift of adjacent slabs is not observed. In the case of the partially protonated hydrated titanate, the Rietveld refinement clearly confirms the relative shift of Nd-containing slabs by half of the a parameter along both horizontal axes accompanied by the space group change from P4/nmm to I4/mmm. In other words, water intercalation results not only in the interlayer space expansion, but also in the conformation change from staggered to eclipsed.
The graphical comparison of the partially protonated hydrated titanate H0.7Na0.3NdTiO4·0.3H2O investigated here and La-containing deuterated compound D0.34Na0.66LaTiO4*0.59D2O reported in [63] is presented in Figure 6. Both structures are practically identical. The distortion of titanium-oxygen octahedra in the form of elongation towards the ion-exchangeable interlayer space direction is hardly noticeable in the La-containing titanate. Equivalent positions of water molecules are significantly distant from each other. In the case of the Nd-containing compound, they are located almost at the same point directly between two sodium cations. In addition, the sodium cation is shifted towards the interlayer space, which is apparently due to the lower occupancy of equivalent water positions in the crystal.
The sodium cations in H0.7Na0.3NdTiO4·0.3H2O are coordinated by nine oxygen atoms, including one belonging to the intercalated water similar to NaNdTiO4 where the ninth oxygen is the axial anion of the opposite layer of the octahedra (Figure 7).
An interesting fact found in this research is a change in the nature of the titanium-oxygen octahedron distortion in the partially protonated phase. The latter is different from NaNdTiO4 and HNdTiO4 in weak elongation of the octahedra towards the ion-exchangeable interlayer space, which allows the Ti4+ cation to be nearly equidistant from both axial oxygen atoms (Figure 8).
Raman spectra of the titanates under consideration are shown in Figure 9. The spectrum of the initial alkaline form NaNdTiO4 is represented by the bands relating to symmetric (897 cm−1) and asymmetric (530, 620 cm−1) stretching modes of axial Ti–O bonds as well as to the vibrations located in the (NdO)2 layer (280–330 cm−1) and in the Na-containing interlayer space (100–200 cm−1). The formation of the partially protonated hydrated derivative H0.7Na0.3NdTiO4·0.3H2O is accompanied by the appearance of characteristic bands referred to the latitudinal vibrations (1620–1630 cm−1) and stretching (2800–3650 cm−1) of the interlayer water. Width and inhomogeneity of the latter point at the unequal involvement of the water molecules in the formation of hydrogen bonds with the titanate. Furthermore, partial protonation leads to the splitting of the symmetric stretching mode of axial Ti–O bonds into two bands (785 and 897 cm−1) confirming the existence of two types of titanium-oxygen octahedra with unequal axial Ti–O distances, whose interlayer vertices are bonded to protons or hydroxonium ions (Ti–OH+ or Ti–OH3O+, the first band) and sodium cations (Ti–ONa+, the second band). The frequencies of the (NdO)2 vibrations are seen to reduce (265 and 305 cm−1) due to the pronounced octahedron distortion. Protonation also results in numerous changes in the low-frequency spectral region (100–200 cm−1) reflecting changes in the interlayer space composition. In the case of the completely protonated compound HNdTiO4, the bands of water vibrations are significantly less intense in comparison with those for H0.7Na0.3NdTiO4·0.3H2O. Thus, HNdTiO4 contains trace water amounts, which is fully consistent with the behavior of its TG curve. The band of the symmetric stretching mode of axial Ti–O bonds occupies a more low-frequency position (835 cm−1) than in the alkaline titanate (897 cm−1), which may be due to the more covalent nature of the OH+ bond and, consequently, greater axial Ti–O distance. At the same time, the frequencies of asymmetric stretching modes of axial Ti–O bonds (545, 610 cm−1) and vibrations of the (NdO)2 layer (280–330 cm−1) prove to be close to those for NaNdTiO4. Consequently, the octahedra in HNdTiO4 are significantly less distorted than in H0.7Na0.3NdTiO4·0.3H2O, which is consistent with the structural analysis data.

4. Conclusions

In the present research, structures of partially and completely protonated Ruddlesden–Popper titanates, namely H0.7Na0.3NdTiO4·0.3H2O and HNdTiO4, have been refined by the Rietveld method on the basis of neutron and X-ray diffraction analysis and compared among themselves as well as with that of the initial alkaline compound NaNdTiO4. It was demonstrated that while interlayer sodium cations in the partially protonated titanate are coordinated by nine oxygen atoms, including one from the intercalated water, in the fully protonated compound the ninth oxygen is the axial anion belonging to the opposite slab of titanium-oxygen octahedra. Furthermore, the partially protonated titanate was revealed to significantly differ from the other two in the octahedron distortion pattern. The compound is characterized by weakly pronounced elongation of the octahedra towards the Nd-containing interlayer space making Ti4+ cations practically equidistant from both axial oxygen atoms, which is accompanied by a low-frequency shift of the bands relating to the asymmetric stretching mode of axial Ti–O bonds observed in the Raman spectra.

Author Contributions

Conceptualization, O.I.S., methodology, O.I.S. and I.A.Z.; investigation, O.I.S., S.A.K. and I.A.M.; data curation, O.I.S.; writing—original draft preparation, O.I.S. and S.A.K.; writing—review and editing, O.I.S. and I.A.R.; visualization, O.I.S. and S.A.K.; supervision, I.A.Z.; funding acquisition, O.I.S. and I.A.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The study was financially supported by Russian Science Foundation (project No. 20-73-00027).

Data Availability Statement

The data presented in this study are available in the article.

Acknowledgments

The study was conducted using the equipment of the National Research Centre «Kurchatov Institute» and the Saint Petersburg State University Research Park: Centre for X-ray Diffraction Studies, Centre for Thermal Analysis and Calorimetry, Centre for Optical and Laser Research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Thangadurai, V.; Shukla, A.; Gopalakrishnan, J. Proton conduction in layered perovskite oxides. Solid State Ion. 1994, 73, 9–14. [Google Scholar] [CrossRef]
  2. Toda, K.; Watanabeb, J.; Satob, M. Synthesis and ionic conductivity of new layered perovskite compound, Ag2La2Ti3O10. Solid State Ion. 1996, 90, 15–19. [Google Scholar] [CrossRef]
  3. Sato, M.; Abo, J.; Jin, T.; Ohta, M. Structure and ionic conductivity of MLaNb2O7 (M-K, Na, Li, H). J. Alloys Compd. 1993, 192, 81–83. [Google Scholar] [CrossRef]
  4. Toda, K.; Suzuki, T.; Sato, M. Synthesis and high ionic conductivity of new layered perovskite compounds, AgLaTa2O7 and AgCa2Ta3O10. Solid State Ion. 1997, 93, 177–181. [Google Scholar] [CrossRef]
  5. Sato, M.; Watanabe, J.; Kazuyoshi, U. Crystal Structure and Ionic Conductivity of a Layered-Perovskite AgLaNb2O7. J. Solid State Chem. 1993, 107, 460–470. [Google Scholar] [CrossRef]
  6. Campbell, K.D. Layered and double perovskites as methane coupling catalysts. Catal. Today 1992, 13, 245–253. [Google Scholar] [CrossRef]
  7. 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]
  8. Hinterding, R.; Feldhoff, A. Two-Dimensional Oxides: Recent Progress in Nanosheets. Z. Phys. Chem. 2018, 233, 117–165. [Google Scholar] [CrossRef]
  9. Hu, Y.; Mao, L.; Guan, X.; Tucker, K.A.; Xie, H.; Wu, X.; Shi, J. Layered perovskite oxides and their derivative nanosheets adopting different modification strategies towards better photocatalytic performance of water splitting. Renew. Sustain. Energy Rev. 2020, 119, 109527. [Google Scholar] [CrossRef]
  10. Tahara, S.; Ichikawa, T.; Kajiwara, G.; Sugahara, Y. Reactivity of the Ruddlesden−Popper Phase H2La2Ti3O10 with Organic Compounds: Intercalation and Grafting Reactions. Chem. Mater. 2007, 19, 2352–2358. [Google Scholar] [CrossRef]
  11. Minich, I.A.; Silyukov, O.I.; Gak, V.V.; Borisov, E.V.; Zvereva, I.A. Synthesis of Organic–Inorganic Hybrids Based on Perovskite-like Bismuth Titanate H2K0.5Bi2.5Ti4 O13·H2O and n-Alkylamines. ACS Omega 2020, 5, 8158–8168. [Google Scholar] [CrossRef] [PubMed]
  12. Silyukov, O.I.; Kurnosenko, S.A.; Zvereva, I.A. Intercalation of Methylamine into the Protonated Forms of Layered Perovskite-Like Oxides HLnTiO4 (Ln = La and Nd). Glas. Phys. Chem. 2018, 44, 428–432. [Google Scholar] [CrossRef]
  13. 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]
  14. Shelyapina, M.G.; Lushpinskaya, I.P.; Kurnosenko, S.A.; Silyukov, O.I.; Zvereva, I.A. Identification of Intercalates and Grafted Organic Derivatives of H2La2Ti3O10 by Multinuclear NMR. Russ. J. Gen. Chem. 2020, 90, 760–761. [Google Scholar] [CrossRef]
  15. Kurnosenko, S.A.; Silyukov, O.I.; Minich, I.A.; Zvereva, I.A. Exfoliation of methylamine and n-butylamine derivatives of layered perovskite-like oxides HLnTiO4 and H2Ln2Ti3O10 (Ln = La, Nd) into nanolayers. Russ. J. Gen. Chem. 2020, 44, 428–432. [Google Scholar]
  16. Rodionov, I.A.; Maksimova, E.A.; Pozhidaev, A.Y.; Kurnosenko, S.A.; Silyukov, O.I.; Zvereva, I.A. Layered Titanate H2Nd2Ti3O10 Intercalated with n-Butylamine: A New Highly Efficient Hybrid Photocatalyst for Hydrogen Production from Aqueous Solutions of Alcohols. Front. Chem. 2019, 7, 863. [Google Scholar] [CrossRef] [PubMed]
  17. Voytovich, V.V.; Kurnosenko, S.A.; Silyukov, O.I.; Rodionov, I.A.; Minich, I.A.; Zvereva, I.A. Study of n-alkylamine Intercalated Layered Perovskite-Like Niobates HCa2Nb3O10 as Photocatalysts for Hydrogen Production from an Aqueous Solution of Methanol. Front. Chem. 2020, 8, 300. [Google Scholar] [CrossRef]
  18. Shelyapina, M.G.; Silyukov, O.I.; Lushpinskaia, I.P.; Kurnosenko, S.A.; Mazur, A.S.; Shenderovich, I.G.; Zvereva, I.A. NMR Study of Intercalates and Grafted Organic Derivatives of H2La2Ti3O10. Molecules 2020, 25, 5229. [Google Scholar] [CrossRef]
  19. Osada, M.; Sasaki, T. Exfoliated oxide nanosheets: New solution to nanoelectronics. J. Mater. Chem. 2009, 19, 2503–2511. [Google Scholar] [CrossRef]
  20. Peláiz-Barranco, A.; González-Abreu, Y. Ferroelectric ceramic materials of the Aurivillius family. J. Adv. Dielectr. 2013, 3, 1330003. [Google Scholar] [CrossRef] [Green Version]
  21. Moure, A. Review and perspectives of Aurivillius structures as a lead-free Piezoelectric system. Appl. Sci. 2018, 8, 62. [Google Scholar] [CrossRef] [Green Version]
  22. Ida, S.; Ogata, C.; Eguchi, M.; Youngblood, W.J.; Mallouk, T.E.; Matsumoto, Y. Photoluminescence of perovskite nanosheets prepared by exfoliation of layered oxides, K2Ln2Ti3O10, KLnNb2O7, and RbLnTa2O7 (Ln: Lanthanide ion). J. Am. Chem. Soc. 2008, 130, 7052–7059. [Google Scholar] [CrossRef] [PubMed]
  23. Bi, D.; Tress, W.; Dar, M.I.; Gao, P.; Luo, J.; Renevier, C.; Schenk, K.; Abate, A.; Giordano, F.; Baena, J.C.; et al. Efficient luminescent solar cells based on tailored mixed-cation perovskites. Sci. Adv. 2016, 2, e1501170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Szczepanski, F.; Bayart, A.; Katelnikovas, A.; Blach, J.F.; Rousseau, J.; Saitzek, S. Luminescence and up-conversion properties in La2Ti2O7:Eu3+,Er3+ oxides under UV and NIR radiations towards a two-color sensor. J. Alloys Compd. 2020, 826, 154157. [Google Scholar] [CrossRef]
  25. Yuan, M.; Dong, W.; Wei, L.; Liu, Q.; Meng, Y.; Wang, X.; Wang, B.; Zhu, B. Stability study of SOFC using layered perovskite oxide La1.85Sr0.15CuO4 mixed with ionic conductor as membrane. Electrochim. Acta 2020, 332, 135487. [Google Scholar] [CrossRef]
  26. Schaak, R.E.; Mallouk, T.E. Perovskites by Design: A Toolbox of Solid-State Reactions. Chem. Mater. 2002, 14, 1455–1471. [Google Scholar] [CrossRef]
  27. Uppuluri, R.; Sen Gupta, A.; Rosas, A.S.; Mallouk, T.E. Soft chemistry of ion-exchangeable layered metal oxides. Chem. Soc. Rev. 2018, 47, 2401–2430. [Google Scholar] [CrossRef] [PubMed]
  28. Dion, M.; Ganne, M.; Tournoux, M. Nouvelles familles de phases MIMII2Nb3O10 a feullets “perovskites”. Mater. Res. Bull. 1981, 16, 1429–1435. [Google Scholar] [CrossRef]
  29. 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] (3 ≤ n ≤ 7). Inorg. Chem. 1985, 24, 3727–3729. [Google Scholar] [CrossRef]
  30. Gopalakrishnan, J.; Bhat, V. A2Ln2Ti3O10 (A = potassium or rubidium; Ln = lanthanum or rare earth): A new series of layered perovskites exhibiting ion exchange. Inorg. Chem. 1987, 26, 4299–4301. [Google Scholar] [CrossRef]
  31. Gopalakrishnan, J.; Sivakumar, T.; Ramesha, K.; Thangadurai, V.; Subbanna, G.N. Transformations of Ruddlesden−Popper Oxides to New Layered Perovskite Oxides by Metathesis Reactions. Chem. Phys. 2000, 122, 6237–6241. [Google Scholar] [CrossRef]
  32. Jacobson, A.J.; Lewandowski, J.T.; Johnson, J.W. Ion exchange of the layered perovskite KCa2Nb3O10 by protons. J. Less Common Met. 1986, 116, 137–146. [Google Scholar] [CrossRef]
  33. Jacobson, A.J.; Johnson, J.W.; Lewandowski, J. Intercalation of the layered solid acid HCa2Nb3O10 by organic amines. Mater. Res. Bull. 1987, 22, 45–51. [Google Scholar] [CrossRef]
  34. Tahara, S.; Sugahara, Y. Interlayer Surface Modification of the Protonated Triple-Layered Perovskite HCa2Nb3O10·xH2O with n-Alcohols. Langmuir 2003, 19, 9473–9478. [Google Scholar] [CrossRef]
  35. Yafarova, L.V.; Silyukov, O.I.; Myshkovskaya, T.D.; Minich, I.A.; Zvereva, I.A. New data on protonation and hydration of perovskite-type layered oxide KCa2Nb3O10. J. Therm. Anal. Calorim. 2020. [Google Scholar] [CrossRef]
  36. Rodionov, I.A.; Silyukov, O.I.; Utkina, T.D.; Chislov, M.V.; Sokolova, Y.P.; Zvereva, I.A. Photocatalytic properties and hydration of perovskite-type layered titanates A2Ln2Ti3O10 (A = Li, Na, K.; Ln = La, Nd). Russ. J. Gen. Chem. 2012, 82, 1191–1196. [Google Scholar] [CrossRef]
  37. Rodionov, I.A.; Mechtaeva, E.V.; Burovikhina, A.A.; Silyukov, O.I.; Toikka, M.A.; Zvereva, I.A. Effect of protonation on the photocatalytic activity of the K2La2Ti3O10 layered oxide in the reaction of hydrogen production. Mon. Chem. Chem. Mon. 2018, 149, 475–482. [Google Scholar] [CrossRef]
  38. Rodionov, I.A.; Fateev, S.A.; Zvereva, I.A. Synthesis of a New Layered Rb2Nd2Ti3O10 Oxide, Its Hydration and Protonation. Glas. Phys. Chem. 2017, 43, 593–596. [Google Scholar] [CrossRef]
  39. Rodionov, I.A.; Sokolova, I.P.; Silyukov, O.I.; Burovikhina, A.A.; Fateev, S.A.; Zvereva, I.A. Protonation and Photocatalytic Activity of the Rb2La2Ti3O10 Layered Oxide in the Reaction of Hydrogen Production. Int. J. Photoenergy 2017, 2017, 9628146. [Google Scholar] [CrossRef] [Green Version]
  40. Silyukov, O.I.; Minich, I.A.; Zvereva, I.A. Synthesis of Protonated Derivatives of Layered Perovskite-Like Bismuth Titanates. Glas. Phys. Chem. 2018, 44, 115–119. [Google Scholar] [CrossRef]
  41. Kurnosenko, S.A.; Silyukov, O.I.; Zvereva, I.A. Preparation of Porous Particles of Layered Perovskite-Like Titanate HLaTiO4. Glas. Phys. Chem. 2020, 46, 272–276. [Google Scholar] [CrossRef]
  42. Byeon, S.; Kileung, P.; Park, K. Structure and Ionic Conductivity of NaLnTiO4, Comparison with Those of Na2Ln2Ti3O10 (Ln = La, Nd, Sm, and Gd). J. Solid State Chem. 1996, 121, 430–436. [Google Scholar] [CrossRef]
  43. Cui, W.; Liu, L.; Ma, S.; Liang, Y.; Zhang, Z. CdS-sensitized K2La2Ti3O10 composite: A new photocatalyst for hydrogen evolution under visible light irradiation. Catal. Today 2013, 207, 44–49. [Google Scholar] [CrossRef]
  44. Takahashi, S.; Nakato, T.; Hayashi, S.; Sugahara, Y.; Kuroda, K. Formation of Methoxy-Modified Interlayer Surface via the Reaction between Methanol and Layered Perovskite HLaNb2O7.cntdot.xH2O. Inorg. Chem. 1995, 34, 5065–5069. [Google Scholar] [CrossRef]
  45. Blasse, G. Crystallographic data of sodium lanthanide titanates (NaLnTiO4). J. Inorg. Nucl. Chem. 1968, 30, 656–658. [Google Scholar] [CrossRef]
  46. Byeon, S.; Yoon, J.J.; Lee, S.O. A New Family of Protonated Oxides HLnTiO4 (Ln = La, Nd, Sm, and Gd). J. Solid State Chem. 1996, 127, 119–122. [Google Scholar] [CrossRef]
  47. Pradhan, D.K.; Samantaray, B.K.; Choudhary, R.N.; Thakur, A.K. Complex impedance studies on a layered perovskite ceramic oxide—NaNdTiO4. Mater. Sci. Eng. B 2005, 116, 7–13. [Google Scholar] [CrossRef]
  48. Pradhan, D.K.; Samantaray, B.K.; Choudhary, R.N.; Thakur, A.K. Complex impedance analysis of NaLaTiO4 electroceramics. J. Mater. Sci. Mater. Electron. 2006, 17, 157–164. [Google Scholar] [CrossRef]
  49. Petrov, A.A.; Melnikova, N.A.; Petrov, A.V.; Silyukov, O.I.; Murin, I.V.; Zvereva, I.A. Experimental investigation and modelling of the Na+ mobility in NaLnTiO4 (Ln = La, Nd) ceramics. Ceram. Int. 2017, 43, 10861–10865. [Google Scholar] [CrossRef]
  50. Toda, K.; Kurita, S.; Sato, M.; Todaa, K.; Kuritab, S.; Satob, M. Synthesis and ionic conductivity of novel layered perovskite compounds, AgLaTiO4 and AgEuTiO4. Solid State Ion. 1995, 81, 267–271. [Google Scholar] [CrossRef]
  51. Toda, K.; Kameo, Y.; Kurita, S.; Sato, M. Crystal structure determination and ionic conductivity of layered perovskite compounds NaLnTiO4 (Ln = rare earth). J. Alloys Compd. 1996, 234, 19–25. [Google Scholar] [CrossRef]
  52. Singh, S.J.; Jayaram, R.V. Chemoselective O-tert-butoxycarbonylation of hydroxy compounds using NaLaTiO4 as a heterogeneous and reusable catalyst. Tetrahedron Lett. 2008, 49, 4249–4251. [Google Scholar] [CrossRef]
  53. Silyukov, O.I.; Abdulaeva, L.D.; Burovikhina, A.A.; Rodionov, I.A.; Zvereva, I.A. Phase transformations during HLnTiO4 (Ln = La, Nd) thermolysis and photocatalytic activity of obtained compounds. J. Solid State Chem. 2015, 226, 101–106. [Google Scholar] [CrossRef]
  54. Reddy, V.; Hwang, D.; Lee, J. Effect of Zr substitution for Ti in KLaTiO4 for photocatalytic water splitting. Catal. Lett. 2003, 90, 39–44. [Google Scholar] [CrossRef]
  55. Kawashima, K.; Hojamberdiev, M.; Chen, S.; Yubuta, K.; Wagata, H.; Domen, K.; Teshima, K. Understanding the effect of partial N3−-to-O2− substitution and H+-to-K+ exchange on photocatalytic water reduction activity of Ruddlesden–Popper layered perovskite KLaTiO4. Mol. Catal. 2017, 432, 250–258. [Google Scholar] [CrossRef]
  56. Rodionov, I.A.; Silyukov, O.I.; Zvereva, I.A. Study of photocatalytic activity of layered oxides: NaNdTiO4, LiNdTiO4, and HNdTiO4 titanates. Russ. J. Gen. Chem. 2012, 82, 635–638. [Google Scholar] [CrossRef]
  57. Berdowski, P.A.; Blasse, G. Luminescence and energy migration in a two-dimensional system: NaEuTiO4. J. Lumin. 1984, 29, 243–260. [Google Scholar] [CrossRef]
  58. Ozawa, T.C.; Ikoshi, A.; Taniguchi, T.; Mizusaki, S.; Nagata, Y.; Noro, Y.; Samata, H.; Takayanagi, S. Low temperature magnetic properties of layered compounds: NaLnTiO4 (Ln = Sm, Eu, Gd, Tb, Dy, Ho and Er). J. Alloys Compd. 2008, 448, 38–43. [Google Scholar] [CrossRef]
  59. Ozawa, T.C.; Ikoshi, A.; Taniguchi, T.; Mizusaki, S.; Nagata, Y.; Noro, Y.; Samata, H.; Takayanagi, S. Magnetic spin interactions observed by heat capacity measurements for layered compounds: NaLnTiO4 (Ln = Sm, Eu, Gd, Tb, Dy, Ho and Er). J. Alloys Compd. 2008, 448, 64–68. [Google Scholar] [CrossRef]
  60. Tezuka, K.; Hinatsu, Y.; Preparation, S. Magnetic Properties of Layered Perovskites NaLnTiO4 (Ln, Sm, Eu, and Gd). J. Solid State Chem. 1998, 346, 342–346. [Google Scholar] [CrossRef]
  61. Silyukov, O.; Chislov, M.; Burovikhina, A.; Utkina, T.; Zvereva, I. Thermogravimetry study of ion exchange and hydration in layered oxide materials. J. Therm. Anal. Calorim. 2012, 110, 187–192. [Google Scholar] [CrossRef]
  62. Zvereva, I.A.; Silyukov, O.I.; Chislov, M.V. Ion-exchange reactions in the structure of perovskite-like layered oxides: I. Protonation of NaNdTiO4 complex oxide. Russ. J. Gen. Chem. 2011, 81, 1434–1441. [Google Scholar] [CrossRef]
  63. Nishimoto, S.; Matsuda, M.; Harjo, S.; Hoshikawa, A.; Ishigaki, T.; Kamiyama, T.; Miyake, M. Neutron diffraction study on protonated and hydrated layered perovskite. J. Solid State Chem. 2006, 179, 3308–3313. [Google Scholar] [CrossRef]
  64. 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]
  65. Abdulaeva, L.D.; Silyukov, O.I.; Petrov, Y.V.; Zvereva, I.A. Low-Temperature Transformations of Protonic Forms of Layered Complex Oxides HLnTiO4 and H2Ln2Ti3O10 (Ln = La, Nd). J. Nanomater. 2013, 2013, 514781. [Google Scholar] [CrossRef] [Green Version]
Figure 1. X-ray diffraction (XRD) patterns of the HNdTiO4, H0.7Na0.3NdTiO4·0.3H2O and NaNdTiO4 samples.
Figure 1. X-ray diffraction (XRD) patterns of the HNdTiO4, H0.7Na0.3NdTiO4·0.3H2O and NaNdTiO4 samples.
Solids 02 00017 g001
Figure 2. Thermogravimetric (TG) curves of partially and completely protonated samples.
Figure 2. Thermogravimetric (TG) curves of partially and completely protonated samples.
Solids 02 00017 g002
Figure 3. Experimental (red), calculated (green) and differential (purple) profiles of the H0.7Na0.3NdTiO4·0.3H2O XRD pattern.
Figure 3. Experimental (red), calculated (green) and differential (purple) profiles of the H0.7Na0.3NdTiO4·0.3H2O XRD pattern.
Solids 02 00017 g003
Figure 4. Experimental (red), calculated (green) and differential (purple) profiles of the HNdTiO4 neutron diffraction pattern.
Figure 4. Experimental (red), calculated (green) and differential (purple) profiles of the HNdTiO4 neutron diffraction pattern.
Solids 02 00017 g004
Figure 5. Comparison of the HNdTiO4 (this work) and DNdTiO4 (data from [64]) structures. The equivalent positions of hydrogen and deuterium are connected by lines.
Figure 5. Comparison of the HNdTiO4 (this work) and DNdTiO4 (data from [64]) structures. The equivalent positions of hydrogen and deuterium are connected by lines.
Solids 02 00017 g005
Figure 6. Comparison of the H0.7Na0.3NdTiO4·0.3H2O (this work) and D0.34Na0.66LaTiO4*0.59D2O (data from [63]) structures. Equivalent positions of water are connected by lines.
Figure 6. Comparison of the H0.7Na0.3NdTiO4·0.3H2O (this work) and D0.34Na0.66LaTiO4*0.59D2O (data from [63]) structures. Equivalent positions of water are connected by lines.
Solids 02 00017 g006
Figure 7. Comparison of the H0.7Na0.3NdTiO4·0.3H2O (this work), NaNdTiO4 (data from [42]) and HNdTiO4 (this work) structures.
Figure 7. Comparison of the H0.7Na0.3NdTiO4·0.3H2O (this work), NaNdTiO4 (data from [42]) and HNdTiO4 (this work) structures.
Solids 02 00017 g007
Figure 8. Titanium-oxygen octahedron geometry in the case of NaNdTiO4 (data from [42]), HNdTiO4 (this work) and H0.7Na0.3NdTiO4·0.3H2O (this work). The titanium-oxygen bond lengths are shown in the figure.
Figure 8. Titanium-oxygen octahedron geometry in the case of NaNdTiO4 (data from [42]), HNdTiO4 (this work) and H0.7Na0.3NdTiO4·0.3H2O (this work). The titanium-oxygen bond lengths are shown in the figure.
Solids 02 00017 g008
Figure 9. Raman spectra of the HNdTiO4, H0.7Na0.3NdTiO4·0.3H2O and NaNdTiO4 samples.
Figure 9. Raman spectra of the HNdTiO4, H0.7Na0.3NdTiO4·0.3H2O and NaNdTiO4 samples.
Solids 02 00017 g009
Table 1. Results of the H0.7Na0.3NdTiO4·0.3H2O structure calculation based on XRD data.
Table 1. Results of the H0.7Na0.3NdTiO4·0.3H2O structure calculation based on XRD data.
H0.7Na0.3NdTiO4·0.3H2O
Tetragonal System, I4/mmm, a = b = 3.7359 Å, c = 27.613 Å, Rwp = 0.1137, R(F2) = 0.0854
IonPositionxyzUisoOccupancy
Na4e000.4225 (25)0.043 (10)0.265 (14)
Nd4e000.30010 (7)0.0006 (25)1
Ti4e000.12489 (17)0.004 (5)1
O14e000.0551 (7)0.0036 (34)1
O28g00.50.13925 (34)0.0084 (26)1
O34e000.2048 (9)0.015 (4)1
Ow8j0.388 (17)0.50.00.0250.314 (14)
Table 2. Results of the HNdTiO4 structure calculation based on neutron diffraction data.
Table 2. Results of the HNdTiO4 structure calculation based on neutron diffraction data.
HNdTiO4
Tetragonal system, P4/nmm a = b = 3.6983 Å, c = 12.0928 Å, Rwp = 0.0253, R(F2) = 0.0627
IonPositionxyzUisoOccupancy
H16e0.434 (12)0.133 (12)0.4966 (25)0.045 (13)0.125
Nd2e0.250.250.8834 (5)0.0001 (13)1
Ti2e0.250.250.2928 (8)0.0152 (35)1
O14g0.250.750.2490 (5)0.0093 (15)1
O22e0.250.250.0759 (6)0.0093 (15)1
O32e0.250.250.4425 (8)0.0177 (23)1
Table 3. Comparison of some bond lengths in HNdTiO4 and DNdTiO4 [64].
Table 3. Comparison of some bond lengths in HNdTiO4 and DNdTiO4 [64].
HNdTiO4DNdTiO4
BondLength, ÅBondLength, Å
H–O3 8x0.919D–O3 8x0.932
Ti–O3 1x1.810Ti–O3 1x1.789
Ti–O1 4x1.923Ti–O1 4x1.919
Ti–O2 1x2.623Ti–O2 1x2.633
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Silyukov, O.I.; Kurnosenko, S.A.; Minich, I.A.; Rodionov, I.A.; Zvereva, I.A. Protonated Forms of Layered Perovskite-Like Titanate NaNdTiO4: Neutron and X-ray Diffraction Structural Analysis. Solids 2021, 2, 265-277. https://0-doi-org.brum.beds.ac.uk/10.3390/solids2030017

AMA Style

Silyukov OI, Kurnosenko SA, Minich IA, Rodionov IA, Zvereva IA. Protonated Forms of Layered Perovskite-Like Titanate NaNdTiO4: Neutron and X-ray Diffraction Structural Analysis. Solids. 2021; 2(3):265-277. https://0-doi-org.brum.beds.ac.uk/10.3390/solids2030017

Chicago/Turabian Style

Silyukov, Oleg I., Sergey A. Kurnosenko, Iana A. Minich, Ivan A. Rodionov, and Irina A. Zvereva. 2021. "Protonated Forms of Layered Perovskite-Like Titanate NaNdTiO4: Neutron and X-ray Diffraction Structural Analysis" Solids 2, no. 3: 265-277. https://0-doi-org.brum.beds.ac.uk/10.3390/solids2030017

Article Metrics

Back to TopTop