Next Article in Journal
Examining Energy Storage Potential in Weakly Polar Nematic Liquid Crystals Infused with Anthraquinone Dye: A Comprehensive Approach
Previous Article in Journal
A Review of Delamination Damage of Composite Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of BaSnO3 as a Highly Dispersed Additive for the Preparation of Proton-Conducting Composites

by
Anton V. Loginov
1,
Alexander I. Aparnev
1,
Nikolai F. Uvarov
1,2,
Valentina G. Ponomareva
2 and
Alexander G. Bannov
1,*
1
Department of Chemistry and Chemical Engineering, Novosibirsk State Technical University, Prospect K. Marksa, 20, 630071 Novosibirsk, Russia
2
Institute of Solid State Chemistry and Mechanochemistry, SB RAS, Kutateladze Str. 18, 630128 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2023, 7(11), 469; https://doi.org/10.3390/jcs7110469
Submission received: 18 October 2023 / Revised: 3 November 2023 / Accepted: 8 November 2023 / Published: 9 November 2023
(This article belongs to the Section Composites Manufacturing and Processing)

Abstract

:
The process of thermolysis of barium hydroxostannate BaSn(OH)6 as a precursor for preparing barium stannate BaSnO3 has been investigated using the method of differential thermal analysis. Thermal decomposition products of the precursor were characterized using X-ray diffraction, IR spectroscopy, low-temperature nitrogen adsorption, and scanning electron microscopy. It was shown that dehydration at nearly 270 °C resulted in the formation of an X-ray amorphous multiphase product, from which single-phase barium stannate crystallized at temperatures above 600 °C. The synthesized barium stannate was used as a functional additive to prepare composite proton electrolytes in the CsHSO4-BaSnO3 system. The structural and transport properties of the obtained system were investigated. It is shown that the highly conductive state of the salt is stabilized in a wide range of temperatures. High conductivity values of composite solid electrolytes in the medium temperature range create the possibility of their use as solid electrolyte membrane materials.

1. Introduction

Recently, materials based on tin dioxide SnO2 doped with various transition and alkaline–earth metals have attracted increased interest and great attention from researchers, which is associated with many promising areas of their application. Among the promising materials based on tin dioxide and stannates are composites and nanocomposites such as MSnO3-SnO2 and M2SnO4-SnO2 (M = Mg, Ca, Ba, Sr, Zn, Cd, Cu, Mn, Co, and Ni). The latter are widely used as components for electronics, optoelectronics, gas sensors, and various catalysts [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27], as well as anode materials for Li-ion batteries and various autonomous devices for energy storage and conversion [28,29,30,31,32,33,34,35,36,37,38,39,40].
Nanomaterials are known to have properties different from those of individual compounds. In particular, nanocomposite solid electrolytes based on acidic salts have enhanced proton conductivity. Among such salts, cesium hydrosulfate exhibits one of the highest conductivity values in the high-temperature phase (10−2 S·cm−1). Cesium hydrosulphate belongs to the solid acids MmHn(XO4)p family (where X = S, Se, P, As; M = Li, Na, K, Rb, Cs, NH4; and n = 1.5, m = 1.5, p = 1.5). These salts undergo a “superprotonic” phase transition at temperatures 50–230 °C accompanied by a sharp increase in the conductivity up to 10−3–10−2 S·cm−1. CsHSO4 in the superprotonic phase is the most conductive salt in this family. However, at the phase transition (414 K), it decreases by four orders of magnitude to σ = 10−6–10−8 S·cm−1. Therefore, it would be desirable to find methods to stabilize the highly conductive state of this compound in the low-temperature region. For this purpose, proton conducting composites based on alkali metal hydrosulfates with various dispersed oxides (1 − x)MHSO4-xA, (M = Cs, Rb, K; A = SiO2, TiO2, Al2O3) were synthesized and studied [41,42,43]. In addition, comprehensive studies of composite systems based on a number of acidic salts of MexHy(AO4)z: MeHSO4 (Me = Cs, Rb, K), Cs5H3(SO4)4, (NH4)3H(SO4)2, CsH2PO4, and silicon dioxide with a different specific surface area (10…580 m2/g) were carried out [44,45,46,47,48].
A number of interesting correlations were found between the concentration of the components, nature, and morphology of the heterogeneous additive and physicochemical properties of the ionic salt in the composite. It was shown that heterogeneous doping of salts with highly dispersed oxide leads to a significant increase in low-temperature conductivity (by orders of magnitude), a decrease in the superionic phase transition temperature, an increase in mechanical strength, and in some cases, the thermal stability of composites. High values of conductivity of composites based on acidic salts and increased mechanical strength give reason to believe that these materials can be promising for use in various kinds of electrochemical devices.
As highly dispersed additives for obtaining proton composite solid electrolytes, it is possible to use stannates of alkali and transition metals MSnO3 (M = Mg, Ca, Sr, Ba, Zn, Ni, etc.), which can be obtained via thermolysis of precursors, -hydroxostannates of the MSn(OH)6 type. It should be noted that stannates obtained via thermolysis are characterized by a relatively high specific surface area (within 10...100 m2/g) and can be used as functional additives for obtaining composite materials [49].
One of the most reliable methods for obtaining compounds of the MSn(OH)6 type (M = Mg, Ca, Sr, Ba, Zn, Ni, etc.) in the form of uniformly mixed powder compositions is the method of co-precipitation of salts. It is based on the precipitation of poorly soluble compounds in the form of hydroxides or metal salts from aqueous solutions of precursor salts by a precipitant solution [50]. The precipitant reacts chemically with the salt solutions, leading to the precipitation of new poorly soluble compounds. If the experiment is set up correctly, a homogeneous mixture of salts with a given ratio of cations can be reproducibly obtained. In the ideal case, the cations from the solution are precipitated simultaneously and at the same rate. The advantages of the chemical precipitation method should be attributed to the granulometric homogeneity of the resulting powders and the high rate of the process. This method avoids the disadvantages inherent in powder technology. For example, it eliminates the need to use the grinding procedure. As a result, the obtained powders are not contaminated by the abrasion products. MSn(OH)6 compounds obtained via precipitation under different conditions [51,52,53,54,55,56,57,58], and consequently their thermal decomposition products, are characterized by a high homogeneity of chemical composition, which is important for their application as functional additives in nanocomposite materials.
Previously, we have shown that the addition of MSn(OH)6 thermolysis products, -nanocomposites based on MSnO3 (M = Mg, Ca, and Sr) with a small admixture of SnO2, leads to an increase in the ionic conductivity of cesium nitrite by more than an order of magnitude [53,59]. Recently, it was found that the addition of zinc stannate to cesium dihydrophosphate leads to an increase in the proton conductivity of the salt in the low-temperature region and an improvement in the stability of the proton electrolyte in the high-temperature region [49]. It is also known that the introduction of oxides with proton conductivity into acidic salts leads to an increase in their proton conductivity [48,60]. Unlike zinc stannate, systems based on barium stannate BaSnO3 are high-temperature proton electrolytes [61,62,63]. Consequently, it can be expected that composites based on acidic salts with the addition of nanocrystalline barium stannate will have increased proton conductivity.
Barium stannate can be obtained via thermolysis of the precursor BaSn(OH)6. It has been found earlier [64,65] that BaSn(OH)6 crystallizes during precipitation as an array of nanorods with a diameter of 20–50 nm and a length up to several micrometers, and the microstructure of its thermolysis products remains unchanged. This makes it possible to use barium stannate as an additive for the creation of composite proton solid electrolytes, for example, on the basis of acidic hydrosulfates. The works on synthesis and investigation of transport properties of proton nanocomposite electrolytes based on acidic salts with the addition of nanocrystalline BaSnO3 have not been carried out before. The aim of the present work is to optimize the conditions for the synthesis of nanocrystalline barium stannate from the BaSn(OH)6 precursor and to study the transport and structural properties of proton composite solid electrolytes CsHSO4-BaSnO3.

2. Materials and Methods

2.1. Materials

The following reagents were used: BaCl2 × 2H2O (GOST 4108-72 “Barium chloride 2-water”, chemical pure, JSC “Reahim”, Donetsk), Na2SnO3 × 3H2O (metastannate 3-water, TU 6-09-1506-76, pure, LLC “Spektr-chem SPb”, St. Petersburg, Russia), and NaOH (GOST 4328-77 “Sodium hydroxide”, chemical pure, LLC “Khimprom, Perm Region, Russia,). CsHSO4 crystals were grown via isothermal evaporation from an aqueous solution of cesium carbonate (99% purity, Rare Metals Plant, Novosibirsk, Russia) and sulfuric acid (pure “ReaKhimLab”, Moscow, Russia) in a stoichiometric ratio. All chemicals were used as received without additional purification. For the preparation of 1M solutions of salts and sodium hydroxide, deionized water obtained using the purification system of the laboratory deionizer BE-2 was used.

2.2. Synthesis of the Precursor BaSn(OH)6

Barium hexahydroxostannate BaSn(OH)6 was synthesized via hydrolytic co-precipitation as follows: First, 0.1 mol of BaCl2 × 2H2O was dissolved in 100 mL of 1M hydrochloric acid solution. To the resulting solution, a 26.67 g (0.1 mol) Na2SnO3 × 3H2O suspension was added, which corresponds to the atomic ratio of Ba:Sn = 1:1. Then, a 1M NaOH solution was gradually added, maintaining the pH of the medium within 8.9...9.4. The pH value was monitored using a laboratory pH meter HI 2221. For complete quantitative co-precipitation of barium and tin (IV), the mixture was stirred continuously for 12 h. The resulting white precipitate was filtered from the mother liquor and washed with distilled water until a negative qualitative reaction to the presence of Cl¯ ions in the solution was achieved and dried in a desiccator at 110 °C for 4 h. The synthesis reaction of BaSn(OH)6 can be written as:
BaCl2 + Na2Sn(OH)6 + 2NaOH = BaSn(OH)6↓ + 2NaCl
The scheme of BaSn(OH)6 precursor preparation is presented in Figure 1.

2.3. Characterization

The microstructure and phase composition of the samples were determined using X-ray diffraction (XRD). X-ray diffraction patterns were recorded at room temperature using a Bruker D8 Advance diffractometer with CuKα radiation in the 2θ range from 10 to 70°. The phases formed in the system were identified using the Crystallographica Search-Match, Version 2.1 program, and the PDF4 database. The average crystallite size was estimated from the diffraction line broadening in X-ray diffraction patterns using the Scherer formula.
d = k λ β c o s θ ,
where d is the average crystallite size, λ is the X-ray wavelength (1.54051 Å), β is the full width at half maximum of the diffraction peak, θ is the diffraction angle, and k = 0.9. Thermal analysis of the dried powders was performed on a NETZSCH Jupiter 449C STA synchronous thermal analyzer coupled to a QMS 403C Aëolos (TG-QMS, Netzsch—Geratebau GmbH, Berlin, Germany) mass spectrometer in an argon flow at temperatures in the range of 20–700 °C at a heating rate of 10 K·min−1. Thermal treatment of the powder was carried out in a SNOL 6.7/1300 muffle furnace at temperatures of 110, 270, 600, and 700 °C for 4 h at each temperature. The microstructure of the samples was studied via field emission scanning electron microscopy (SEM) on a Hitachi SU8000 electron microscope (Hitachi, Tokyo, Japan). The images were taken in the secondary electron mode at an accelerating voltage of 2–30 kV and a working distance of 8–10 mm. The sample’s electron dispersion spectra (EDS) were analyzed using an Oxford Instruments X-max energy dispersive spectrometer. Analytical measurements of the EDS were optimized using a previously established method, as outlined in the references [66]. Prior to the measurements, the samples were affixed to an aluminum mount with a 25 mm diameter and secured firmly with conductive graphite adhesive tape. Additionally, the morphology of the unmodified samples was studied to preclude any surface effects due to the application of the conductive layer, as noted in [67]. The specific surface area values and the average pore size were calculated using the Brunauer–Emmett–Teller (BET) method described in detail in papers [68,69]. Infrared spectra were recorded on a Carry 660 FTIR (Fourier transform infrared) spectrometer (Agilent Technologies, Santa Clara, CA, USA) with a PIKE Technologies Gladi ATR (diamond crystal) broken total internal reflection attachment in the range of 500–4000 cm−1. The samples were prepared as vacuum-pressed KBr pellets with an admixture of the compound under study. The composites in the system (1 − x)CsHSO4-xBaSnO3 (where x = 0.2, 0.4-mol fraction of barium stannate) were synthesized by thoroughly mixing the components in an agate mortar and heating the granulated samples for 10–20 min at a temperature of 200–210 °C, close to the melting point of CsHSO4. The proton conductivity of the composites was measured using a two-electrode scheme on an alternating current using an Instek LCR-821 Impedance Meter (frequency range 12 Hz–200 kHz) and a precision electrochemical meter IPU-1RLK-1/2008 (CIT LLC, Moscow, Russia) (frequency range 1 Hz–3.3 MHz). To measure the electrical conductivity, pellets with a diameter of 6 mm and a thickness of 1–1.5 mm were pressed from the samples under a pressure of 30–50 MPa. The electrical conductivity of the samples with silver or platinum electrodes was measured under cooling at a rate of 1–2 °C/min in an air atmosphere at a relative humidity of RH = 30%.

3. Results and Discussion

3.1. Synthesis of Precursor BaSn(OH)6

The X-ray diffraction patterns of the sample obtained as a result of co-deposition and drying at 110 °C in air (Figure 2) show reflexes characteristic of barium hexahydroxostannate BaSn(OH)6, which has a crystal structure with a monoclinic unit cell (space group symmetry P21/n, PDF4, card no. 09-0053). Along with the BaSn(OH)6 phase, the diffractogram of the sample shows reflexes related to the barium carbonate BaCO3 phase with an orthorhombic unit cell of the witherite type (symmetry space group Pmcn, PDF4, card no. 05-00378).
The processing of diffractograms using the Rietveld method in the Topas v. 4.2 program allowed us to determine the unit cell parameters of BaSn(OH)6 and BaCO3 phases in the initial sample. The unit cell parameters of BaSn(OH)6 (a = 9.3892 ± 0.0004 Å, b = 6.3400 ± 0.0003 Å, c = 10.5649 ± 0.0005 Å, β = 113.211 ± 0.003°, α = γ = 90°, Z = 4, Vcell = 577.96 Å3, and ρ = 4.115 g/cm3) are close to the corresponding values for this compound reported earlier [64]. A small difference in the values of the parameter a and the angle β is observed. The best-fitting parameters are obtained under the assumption that BaSn(OH)6 crystallites have a size of about 175 nm and a preferred (200) orientation. The BaCO3 unit cell parameters (a = 5.205 ± 0.002 Å, b = 9.120 ± 0.004 Å, c = 6.430 ± 0.002 Å, Vcell = 305.22 Å3, and ρ = 4.294 g/cm3) are consistent with the data presented in the PDF database (PDF4, card no. 05-00378), and the particle size of the witherite phase is 28 nm. Analysis of diffraction data using the Rietveld method shows that the synthesized BaSn(OH)6 sample contains 33 weight % BaCO3. The presence of witherite impurity can be explained by the reaction of BaSn(OH)6 with carbon dioxide contained in the air:
BaSn(OH)6 + CO2 = BaCO3 + SnO2·xH2O + (3 − x)H2O
Nanocrystalline BaCO3 and an equivalent amount of amorphous tin dioxide hydrate SnO2 × xH2O, which does not give reflexes on X-ray diffractograms, are formed. According to the data of scanning electron microscopy (Figure 3a), a freshly deposited mixed hydroxide BaSn(OH)6 sample consists of clearly visible rod-shaped particles 10–50 microns in size, the surface of which is covered with a loose layer, apparently consisting of a mixture of nanosized particles of barium carbonate and hydrated tin dioxide. According to the results of chemical microanalysis carried out using energy dispersive spectroscopy, the atomic ratio Ba:Sn:O lies in the range (10 ± 1):(12 ± 2):(65 ± 3), which is close to the stoichiometric ratio for both BaSn(OH)6 and the mixture BaCO3–SnO2 × xH2O (at a molar ratio of components 1:1). It was found that the amount of barium carbonate in the sample increases when the freshly deposited BaSn(OH)6 precursor is stored in air (Figure 3b).

3.2. Characteristics of Intermediate Products of Thermolysis and Final Sample of Barium Stannate

The processes occurring during the heating of the freshly deposited BaSn(OH)6 sample were studied using the method of synchronous thermal analysis; the results obtained are presented in Figure 4. The process of BaSn(OH)6 thermolysis includes two stages of dehydration, which correspond to endothermic effects: a weak one with a maximum at 160 °C and a strong effect at Tmax = 270 °C. The dehydration reaction is indicated by a change in the ionic current of the mass spectrometric sensor corresponding to the release of water molecules with an atomic mass of 18 amu. The dehydration process is fully completed at a temperature of nearly 500 °C and can be described by the equations:
BaSn(OH)6 → BaSnO3 + SnO2 + 3H2O
SnO2 × xH2O = SnO2 + xH2O
Taking into account that the initial sample contains BaSn(OH)6, BaCO3 (being in weight ratio 0.67:0.33), and amorphous phase SnO2 × xH2O, the molar fraction of which is equal to the molar fraction of BaCO3, it is possible to calculate the residual value of relative mass after full dehydration. These values are 86.9 wt.% and 89.8 wt.% for the limiting values of the degree of hydration of tin dioxide x = 2 (tin acid) and x = 0 (anhydrous tin dioxide), respectively. The experimentally found value of the residual mass fraction is 88 wt.%, which corresponds to the value x = 1.24 or the calculated formula of hydrated tin dioxide SnO2 × 1.24H2O.
Upon further heating of the sample in the temperature range of 500–650 °C, an exothermic peak is observed. The thermal effect is accompanied by a symbatic release of carbon dioxide molecules with m/z of 44 amu, which is recorded using a mass spectrometric sensor (Figure 4), which suggests that the thermal effect is caused by the occurrence of the chemical reaction
BaCO3 + SnO2 → BaSnO3 + CO2,
in which barium carbonate, formed as a result of reaction (3), reacts with tin dioxide as the product of the amorphous SnO2 × xH2O dehydration. Taking into account the weight fraction of barium carbonate in the initial mixture (33 wt.%) and the equality of molar fractions of BaCO3 and SnO2, it is possible to calculate the limiting value of the relative mass of the product after the dehydration reactions (4) and (5) and CO2 elimination (6). The calculated value, 84.9 wt.%, is close to the experimental value of 85.3 wt.%, which indicates the adequacy of the interpretation of the thermogravimetric analysis data.
The results of diffraction studies, shown in Figure 5, agree well with the data of thermal analysis. On the diffractogram obtained after heating the sample at 500 °C, i.e., after complete dehydration of the sample, there are only reflections related to the BaCO3 witherite phase, which was initially contained in the sample (Figure 5, curve 2). Consequently, the dehydration products, BaSnO3 and SnO2, formed as a result of reactions (4) and (5) at this temperature are X-ray amorphous.
After heating the sample at 600 °C, the reflections related to the crystalline BaSnO3 phase appear on the diffractogram, and the intensity of the diffraction peaks of the witherite BaCO3 phase decreases significantly. According to the data of thermal analysis at this temperature, the exothermic effect and chemical reaction (5) with the formation of CO2 are observed. These processes are associated with the crystallization process, the transition of the amorphous phase of barium stannate into the crystalline phase, and the formation of a nanocomposite, which is a mixture of BaSnO3 with residual unreacted nanocrystalline BaCO3 (Figure 5, curve 3). At further increase of the heating temperature up to 700 °C, the reaction (6) comes to an end, and the diffractogram of the sample (Figure 5, curve 4) corresponds to monophasic barium stannate with the structure of perovskite with a cubic unit cell (symmetry space group Pm3m). The lattice parameter of the obtained sample a = 4.1139 ± 0.0002 Å agrees with the literature data (PDF4, card no. 015-00780 and 03-00675) [64,70,71].
A comparison of electron microscopic images of the sample heated at 700 °C (Figure 6) and the original sample (Figure 3) shows that heating the sample leads to an unusual change in its morphology. Initially, aggregated particles covered with a rough film turn into an array of nanorods with a thickness of 10–50 nm and a length of a few microns after heating. This change in morphology can be explained by two processes: the formation of pseudomorphosis with a preferred orientation of particles during the dehydration of the initial BaSn(OH)6 crystallites and the subsequent selective crystallization of the amorphous phases on the end surfaces of the pseudomorphosis. As a result, the impurity phases BaCO3 and SnO2⋅xH2O localized on the surface of crystallites are removed from most of the surface and recrystallize in the end region of the pseudomorphosis. Finally, the particles of the pseudomorphosis become thinner and longer, forming nanorods.
The analysis of FT spectra of the synthesized compounds allows us to confirm their composition and some features of their structure. The characteristic bands in the spectra were identified based on the literature data [72,73]. The FTIR spectra of the obtained samples of barium hydroxanthanate BaSn(OH)6 (Figure 7, curve 1) and barium stannate BaSnO3 (Figure 7, curve 2) agree with those obtained earlier [74]. The spectrum of BaSn(OH)6 clearly shows a broad band in the region of ~562 cm−1 corresponding to the vibrations of Sn-O bonds. The absorption band in the frequency region of ~750 cm−1 is characteristic of bonds of more than one type of Ba-O and Sn-O-Ba groups. A series of bands in the 875–980 cm−1 region corresponds to deformation vibrations of Sn-OH groups having partially ionic character [73,75]. The appearance of bands at ~1475 and 1254 cm−1 indicates the presence of carbonate impurities in the sample, and the presence of these peaks at different frequencies can be explained by the conjugate motion of C-O and C=O bonds. Carbonates, mainly BaCO3, can also be formed as a result of the absorption of carbon dioxide from the atmosphere directly during FTIR spectra [76]. The broad absorption of medium intensity at 1630–1750 cm−1 definitely indicates the presence of water, while the broad band of valence vibrations ν(OH) in the region at ~3410 cm−1 shows the presence of hydroxyl groups in the composition of the obtained BaSn(OH)6 precursor (Figure 7, curve 1).
In the FTIR spectrum of BaSnO3 (Figure 7, curve 2), a strong band at 715 cm−1 is observed, related to the symmetric valence vibrations of Sn-O bonds. The presence of physically adsorbed moisture (δ(OH) strain vibrations) is also recorded above 1630 cm−1. These bands decrease their intensity as a result of the dehydration of the sample during heat treatment at 600 °C, as evidenced by the mass loss from thermogravimetry results. It was found that as a result of exposure of the BaSnO3 sample to air, a broad diffuse absorption band in the region of ~3410 cm−1 corresponding to the valence vibrations of ν(H2O) is again recorded in the IR spectrum.
Table 1 shows the results of studies of the textural properties of the obtained samples calcined at different temperatures for 4 h. It can be seen that the pseudomorphosis formed during the thermolysis of BaSn(OH)6 is characterized by high values of specific surface area and the presence of mesopores with sizes in the range of 2–6 nm. With increasing heating temperature, the values of specific surface area monotonically decrease.
The decrease in the specific surface area of composites during calcination is associated with the growth of the grain size. The crystallite size L can be estimated from the known values of specific surface area S′ using a simplified expression obtained for spherical or cubic particles:
L = 6 ρ × S ,
where ρ is the density of the substance (for BaSnO3 the value of ρ = 7.236 g·cm−3). The evaluation shows that by increasing the heating temperature from 600 °C to 700 °C, the size of BaSnO3 particles increases from 36 to 55 nm, which is close to the corresponding values estimated from the broadening of diffraction peaks on diffractograms. Further, the sample of barium stannate obtained by heating the initial sample at 700 °C was used as an oxide additive for the preparation of proton composite solid electrolytes.

3.3. Study of Transport Properties of Composites

Cesium hydrosulphate exists in a low-temperature (LT) monoclinic phase (spacegroup P21/m) at room temperature, which contains statistically symmetric hydrogen bonds organized in infinite (HSO4−)n chains. At 141 °C, CsHSO4 transforms to the high-temperature superprononic tetragonal phase (I41/amd) with a conductivity change from 10−6 up to 3 × 10−2 S·cm−1 and proton transfer numbers equal to one [77]. Figure 8 shows temperature dependences of the conductivity of pure CsHSO4 and composites (1 − x)CsHSO4-xBaSnO3 of different compositions (where x-mole fraction). As seen, at low temperatures, the introduction of a heterogeneous additive leads to a significant (up to 2.5 orders of magnitude) conductivity rise with an increase in the BaSnO3 mole fraction up to x = 0.2 (13.2 vol.%). The conductivity enhancement depends markedly on the BaSnO3 additive concentration.
Accordingly, the magnitude of the conductivity jump at the phase transition decreases with increasing x; the conductivity jump due to the phase transition becomes much smoother and shifts to the temperature region of ~85 °C. The shift of the phase transition temperature indicates the stabilization of the high-temperature phase in the low-temperature region. The proton conductivity of the composite 0.8CsHSO4-0.2BaSnO3 is 3 × 10−3 S/cm−1 at 160 °C. With further growth of the additive fraction up to x = 0.4 (28.3 vol.% of BaSnO3), the conductivity decreases both in the high-temperature and low-temperature regions. The activation energy of conductivity decreases to 0.47 eV. The superionic phase transition completely disappears, and there is only a slight change in the slope at 85 °C on the temperature dependence. The conductivity of the high-temperature phase decreases with increasing additive fraction by an order of magnitude at x = 0.2 and by about 1.5 orders of magnitude at x = 0.4 (see Figure 8). The observed composite effect is similar to (1 − x)CsHSO4-xSiO2 (or TiO2) systems; however, the conductivity enhancement is somewhat different [41,42,43].
X-ray diffraction data (Figure 9) show that the diffraction pattern of the composite may be represented as a sum of contributions of individual phases. The structure of the CsHSO4 salt is preserved in the composite systems with a significant decrease in the intensity of reflections and their significant broadening, which may be caused by the dispersion and partial amorphization of the salt in (1 − x)CsHSO4-xBaSnO3 composites due to the surface interaction between acid salt and the additive. With the increase in the additive fraction in the composites, the X-ray diffraction contribution of BaSnO3 (marked with an asterisk) becomes more pronounced.

4. Conclusions

In this work, we have demonstrated the possibility of the preparation of nanocrystalline barium stannate BaSnO3 via the thermal decomposition of barium hexahydroxostannate precursor BaSn(OH)6. The latter, characterized by rod morphology up to 10–50 μm in size, was synthesized via hydrolytic precipitation from solutions of barium chloride BaCl2 and sodium stannate Na2SnO3. It was shown that thermolysis at a temperature of about 270 °C resulted in the formation of an X-ray amorphous multiphase product consisting of amorphous phases of hydrated tin dioxide and barium stannate, as well as an impurity of crystalline barium carbonate BaCO3. During thermal treatment of the sample at 600 °C, barium stannate (with a specific surface area of 23 m2/g) with traces of barium carbonate is formed. At temperatures of 600–700 °C, the BaSnO3 phase crystallizes with a particle size of 36–55 nm. As a result of the study, it is shown that heating at temperatures of 600–700 °C is the optimal condition for obtaining single phase and highly dispersed barium stannate.
The proton conductivity of composite systems of composition (1 − x)CsHSO4-xBaSnO3 (x = 0.2–0.4) has been investigated. The crystal structure of CsHSO4 is preserved in the composite systems with a significant decrease in the intensity of reflexes and their significant broadening due to dispersion and partial amorphization of the salt in (1 − x)CsHSO4-xBaSnO3. As a result of the conductivity study, it is shown that the introduction of the heterophase BaSnO3 additive leads to a significant (up to 2.5 orders of magnitude) increase in conductivity in the low-temperature region with an increasing mole fraction of BaSnO3 up to x = 0.2. An increase in the conductivity is accompanied by the shift of the superionic phase transition to a lower temperature region and its practical disappearance at x = 0.4. The proton conductivity of the composite 0.8CsHSO4-0.2BaSnO3 reaches 3 × 10−3 S·cm−1 at 160 °C. This will allow the use of this material in electrochemical applications. Thus, it is demonstrated that highly dispersed barium stannate BaSnO3 is a suitable heterogeneous additive in composite solid electrolytes. CsHSO4-BaSnO3 composite solid electrolytes can be used as solid electrolyte membrane materials for hydrogen production in the medium-temperature region. This approach may prove to be competitive with alternative technologies.

Author Contributions

A.V.L. and A.G.B.: investigation, methodology, data curation, and writing—original draft. A.I.A., A.V.L., N.F.U. and V.G.P.: investigation, data curation, and methodology. N.F.U.: conceptualization, paper revision, and English correction. V.G.P.: writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by the Russian Federation Ministry of Science and Higher Education (state research target, project no. FSUN-2023-0008).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, D.H.; Qu, L.; Kim, D.W.; Kang, S.B.; Park, J.-S.; Cho, I.S. Role of cations in the photovoltaic performance optimization of ternary stannates, MSnO3 (M = Ca, Sr, and Ba) and N2SnO4 (N = Ca, Sr, Ba, and Zn). Ceram. Int. 2023, 49, 32015–32023. [Google Scholar] [CrossRef]
  2. Chahib, S.; Leroy, G.; Duponchel, B.; Poupin, C.; Ez-zahraouy, H.; Fasquelle, D. Investigation of structural, morphological, and dielectric properties of BaSnO3 ceramics and thin films prepared by sol-gel method. Ceram. Int. 2023, 49, 17542–17553. [Google Scholar] [CrossRef]
  3. Muraleedharan, S.; Ashok, A.M. Nitrogen doping for effective enhancement of optoelectronic performance of ASnO3 (A-Ca, Sr, Ba) transparent conducting oxides deposited by doctor blade method. Optik 2023, 287, 171096. [Google Scholar] [CrossRef]
  4. Khatun, M.; Mitra, P.; Mukherjee, S. Effect of band gap and particle size on photocatalytic degradation of NiSnO3 nanopowder for some conventional organic dyes. Hybrid Adv. 2023, 4, 100079. [Google Scholar] [CrossRef]
  5. Gao, Y.; Wang, X.; Zhang, Q.; Wang, H.; Xu, G.; Wang, X. Influence of La doping on the ethanol gas sensing properties of CdSnO3 micro-cubes. Sens. Actuators B 2023, 394, 134447. [Google Scholar] [CrossRef]
  6. Marikutsa, A.; Dobrovolskii, A.A.; Rumyantseva, M.N.; Mikhaylov, A.A.; Medvedev, A.G.; Lev, O.; Prikhodchenko, P.V. Improved H2S sensitivity of nanosized BaSnO3 obtained by hydrogen peroxide assisted sol-gel processing. J. Alloys Compd. 2023, 944, 169141. [Google Scholar] [CrossRef]
  7. Venkatesh, G.; Suganesh, R.; Jayaprakash, J.; Srinivasan, M.; Prabu, K.M. Perovskite type BaSnO3-reduced graphene oxide nanocomposite for photocatalytic decolourization of organic dye pollutant. Chem. Phys. Lett. 2022, 787, 139237. [Google Scholar] [CrossRef]
  8. Bogdan, T.V.; Krasnikov, P.A.; Smirnov, A.V.; Koklin, A.E.; Mashchenko, N.V.; Bogdan, V.I. Utilization of Acetone, by-product of cumene process for phenol production, via BaSnO3-catalyzed aldol condensation. Dokl. Phys. Chem. 2022, 507, 147–152. [Google Scholar] [CrossRef]
  9. Salem, F.Z.; Ahmed, M.A.; Sadek, M.A.; Elmahgary, M.G. Novel hydrogen-doped SrSnO3 perovskite with excellent optoelectronic properties as a potential photocatalyst for water splitting. Int. J. Hydrogen Energy 2022, 47, 18321–18333. [Google Scholar] [CrossRef]
  10. Ashiq, M.G.B.; Mahmood, Q.; Haq, B.U.; Flemban, T.H.; Kattan, N.A.; Alshahrani, T.; Laref, A. The study of electronics, optoelectronics, thermoelectric, and mechanical properties of Zn/CdSnO3 perovskites. Mater. Sci. Semicond. Process. 2022, 137, 106229. [Google Scholar] [CrossRef]
  11. Li, X.; Gao, X.; Lu, W.; Fouad, H.; Mourad, A.-H.I.; Akhtar, M.S.; Guo, W. Hydrothermal synthesis of hollow CoSnO3 nanocubes for highly response and selective ethanol gas sensing. Mater. Lett. 2022, 316, 132056. [Google Scholar] [CrossRef]
  12. Sanchela, A.V.; Wei, M.; Cho, H.J.; Ohta, H. Optoelectronic properties of transparent oxide semiconductor ASnO3 (A = Ba, Sr, and Ca) epitaxial films and thin film transistors. J. Sci. Technol. A 2022, 40, 020803. [Google Scholar] [CrossRef]
  13. Qi, C.; Zhang, C.; Yang, Z. Constructing heterointerface of crystalline Au nanoparticles and amorphous porous CoSnO3 nanocubes for sensitive electrochemical detection of glucose. Microchem. J. 2022, 183, 108039. [Google Scholar] [CrossRef]
  14. Tao, F.; Li, F.; Huang, J.; Xue, Z.; Yu, C.; Cai, Z.; Pei, L. A General Hydrothermal Growth and Photocatalytic Performance of Barium Tin Hydroxide/Tin Dioxide Nanorods. Cryst. Res. Technol. 2021, 57, 2100156. [Google Scholar] [CrossRef]
  15. Sá, B.S.; Zito, C.A.; Perfecto, T.M.; Volanti, D.P. Porous ZnSnO3 nanocubes as a triethylamine sensor. Sens. Actuators B 2021, 338, 129869. [Google Scholar] [CrossRef]
  16. Ochoa-Munoz, Y.H.; Rodriguez-Paez, J.E.; de Gutierrez, R.M. Structural and optical study of perovskite nanoparticles MSnO3 (M = Ba, Zn, Ca) obtained by a wet chemical route. Mater. Chem. Phys. 2021, 266, 124557. [Google Scholar] [CrossRef]
  17. Wang, J.; Luo, B. Optical and electronic properties in amorphous BaSnO3 thin films. Phys. B Condens. Matter. 2021, 601, 412586. [Google Scholar] [CrossRef]
  18. Kramer, J.W.; Kelly, B.; Manivannan, V. Synthesis of MSn(OH)6 (where M = Mg, Ca, Zn, Mn, or Cu) materials at room temperature. Cent. Eur. J. Chem. 2010, 8, 65–69. [Google Scholar] [CrossRef]
  19. Chu, X.; Gan, Z.; Bai, L.; Dong, Y.; Rumyantseva, M.N. The acetic acid vapor sensing properties of BaSnO3 microtubes prepared by electrospinning method. Mater. Sci. Eng. B 2020, 259, 114606. [Google Scholar] [CrossRef]
  20. Arora, I.; Praveen, K. Effect of annealing temperature on structure-property correlations in Zn2SnO4 nanostructured films for optoelectronics. Mater. Res. Express 2020, 7, 035023. [Google Scholar] [CrossRef]
  21. Debata, S.; Banerjee, S.; Chakraborty, S.; Sharma, P.K. Template assisted hydrothermal synthesis of CoSnO3 hollow microspheres for electrocatalytic oxygen evolution reaction. Int. J. Hydrogen Energy 2019, 44, 21623–21636. [Google Scholar] [CrossRef]
  22. Gomez-Solis, C.; Oliva, J.; Diaz-Torres, L.A.; Bernal-Alvarado, J.; Reyes-Zamudio, V.; Abidov, A.; Torres-Martinez, L.M. Efficient photocatalytic activity of MSnO3 (M: Ca, Ba, Sr) stannates for photoreduction of 4-nitrophenol and hydrogen production under UV light irradiation. J. Photochem. Photobiol. A Chem. 2019, 371, 365–373. [Google Scholar] [CrossRef]
  23. Rembeza, S.I.; Belousov, S.A.; Kosheleva, N.N.; Rembeza, E.S.; Svistova, T.V.; Suvaci, E.; Özel, E.; Tuncolu, G.; Açiksari, C. Amorphous films of ternary zinc and tin oxides for transparent electronics. Tech. Phys. Lett. 2018, 44, 984–987. [Google Scholar] [CrossRef]
  24. Sahmi, A.; Laib, R.; Omeiri, S.; Bensadok, K.; Trari, M. Photoelectrochemical properties of the perovskite BaSnO3 synthetized by chemical route. Application to electro-photocatalytic mineralization of ibuprofen. J. Photochem. Photobiol. A Chem. 2018, 364, 443–448. [Google Scholar] [CrossRef]
  25. Honorio, L.M.C.; Santos, M.V.B.; Da Silva Filho, E.C.; Osajima, J.A.; Maia, A.S.; Santos, I.M.G. Alkaline Earth Stannates Applied in Photocatalysis: Prospection and Review of Literature. Ceram. 2018, 64, 559–569. [Google Scholar] [CrossRef]
  26. Inagaki, M.; Kuroishi, T.; Yamashita, Y.; Urata, M. Syntheses of MSn(OH)6 by coprecipitation and of MSnO3 by thermal decomposition (M = Mg, Co, Zn, Mn, Cd, Ca, Sr, Ba). Z. Anorg. Allg. Chem. 1985, 527, 193–202. [Google Scholar] [CrossRef]
  27. Jayathilaka, W.A.D.M.; Chinnappan, A.; Ji, D.; Ghosh, R.; Tran, T.Q.; Ramakrishna, S. Facile and Scalable Electrospun Nanofiber-Based Alternative Current Electroluminescence (ACEL) Device. ACS Appl. Electron. Mater. 2021, 3, 267–276. [Google Scholar] [CrossRef]
  28. Kohan, M.; Mahmoudi, T.; Wang, Y.; Im, Y.H.; Hahn, Y.-B. SnO2/BaSnO3 electron transport materials for stable and efficient perovskite solar cells. Appl. Surf. Sci. 2023, 613, 156068. [Google Scholar] [CrossRef]
  29. Zvonareva, I.A.; Medvedev, D.A. Proton-conducting barium stannate for high-temperature purposes: A brief review. J. Eur. Ceram. Soc. 2023, 43, 198–207. [Google Scholar] [CrossRef]
  30. Antonio, J.E.; Rosas-Huerta, J.L.; Cervantes, J.M.; León-Flores, J.; Romero, M.; Carvajal, E.; Escamilla, R. Li/Na atoms’ substitution effects on the structural, electronic, and mechanical properties of the CaSnO3 perovskite for battery applications. Comp. Mater. Sci. 2023, 219, 112006. [Google Scholar] [CrossRef]
  31. Guo, S.; Liu, M.; Yang, H.; Feng, X.; Bai, Y.; Wu, C. CoSnO3/C nanocubes with oxygen vacancy as high-capacity cathode materials for rechargeable aluminum batteries. Green Energy Environ. 2023, 8, 883–892. [Google Scholar] [CrossRef]
  32. Roy, A.; Ghosh, P.; Devi, P.S.; Sundaram, S.; Mallick, T.K. Efficient carbon counter electrodes for BaSnO3-based dye-sensitized solar cells. Mater. Today Proc. 2021, 45, 3685–3691. [Google Scholar] [CrossRef]
  33. Wei, J.L.; Jin, X.Y.; Yu, M.C.; Wang, L.; Guo, Y.H.; Dong, S.T.; Zhang, Y.M. Electrospun ZnSnO3/C nanofibers as an anode material for lithium-ion batteries. J. Electron. Mater. 2021, 50, 4945–4953. [Google Scholar] [CrossRef]
  34. Sim, C.K.; Majid, S.R.; Mahmood, N.Z. ZnSnO3/mesoporous biocarbon composite towards sustainable electrode material for energy storage device. Microchem. J. 2021, 164, 105968. [Google Scholar] [CrossRef]
  35. Purushothamreddy, N.; Kovendhan, M.; Dileep, R.K.; Veerappan, G.; Kumar, S.K.; Joseph, D.P. Synthesis and characterization of nanostructured La-doped BaSnO3 for dye-sensitized solar cell application. Mater. Chem. Phys. 2020, 250, 123137. [Google Scholar] [CrossRef]
  36. Roy, A.; Das, P.P.; Selvaraj, P.; Devi, P.S.; Sundaram, S. Template free synthesis of CdSnO3 micro-cuboids for dye sensitized solar cells. J. Photochem. Photob. A Chem. 2019, 380, 111824. [Google Scholar] [CrossRef]
  37. Cheng, H.; Koh, K.L.P.; Liu, P.; Thang, T.Q.; Duong, H.M. Continuous self-assembly of carbon nanotube thin films and their composites for supercapacitors. Colloids Surf. A Physicochem. Eng. Aspects. 2015, 481, 626–632. [Google Scholar] [CrossRef]
  38. Veerappan, G.; Yoo, S.; Zhang, K.; Ma, M.; Kang, B.; Park, J.H. High-reversible capacity of Perovskite BaSnO3/rGO composite for Lithium-Ion Battery Anodes. Electrochim. Acta 2016, 214, 31–37. [Google Scholar] [CrossRef]
  39. Jena, H.; Kutty, K.V.G.; Kutty, T.R.N. Ionic transport and structural investigations on MSn(OH)6 (M = Ba, Ca, Mg, Co, Zn, Fe, Mn) hydroxide perovskites synthesized by wet sonochemical methods. Mater. Chem. Phys. 2004, 88, 167–179. [Google Scholar] [CrossRef]
  40. Wang, Y.; Li, D.; Liu, Y.; Zhang, J. Fabrication of novel rugby-like ZnSnO3/reduced graphene oxide composites as a high-performance anode material for lithium-ion batteries. Mater. Lett. 2016, 167, 222–225. [Google Scholar] [CrossRef]
  41. Ponomareva, V.G.; Uvarov, N.F.; Lavrova, G.V.; Hairetdinov, E.F. Composite protonic solid electrolytes in the CsHSO4-SiO2 system. Solid State Ionics 1996, 90, 161–167. [Google Scholar] [CrossRef]
  42. Ponomareva, V.G.; Lavrova, G.V. Influence of dispersed TiO2 on protonic conductivity of CsHSO4. Solid State Ionics 1998, 106, 137–141. [Google Scholar] [CrossRef]
  43. Ponomareva, V.G.; Lavrova, G.V.; Simonova, L.G. The influence of heterogeneous dopant porous structure on the properties of protonic solid electrolyte in the CsHSO4-SiO2 system. Solid State Ionics 1999, 118, 317–323. [Google Scholar] [CrossRef]
  44. Ponomareva, V.G.; Lavrova, G.V. The investigation of disordered phases in nanocomposite proton electrolytes based on MeHSO4 (Me = Rb, Cs, K). Solid State Ionics 2001, 145, 197–204. [Google Scholar] [CrossRef]
  45. Mohammad, N.; Mohamad, A.B.; Kadhum, A.A.H.; Shyuan, L.K. Conductivity and thermal stability of solid acid composites CsH2PO4/NaH2PO4/SiO2. Malays. J. Analytic. Sci. 2016, 20, 633–641. [Google Scholar] [CrossRef]
  46. Navarrete, L.; Yoo, C.-Y.; Serra, J.M. Comparative Study of Epoxy-CsH2PO4 Composite Electrolytes and Porous Metal Based Electrocatalysts for Solid Acid Electrochemical Cells. Membranes 2021, 11, 196. [Google Scholar] [CrossRef] [PubMed]
  47. Ponomareva, V.G.; Lavrova, G.V. Controlling the proton transport properties of solid acids via structural and microstructural modification. J. Solid State Electrochem. 2011, 15, 213–221. [Google Scholar] [CrossRef]
  48. Anfimova, T.; Jensen, A.H.; Christensen, E.; Jensen, J.O.; Bjerrum, N.J.; Li, Q. CsH2PO4/NdPO4 composites as proton conducting electrolytes for intermediate temperature fuel cells. J. Electrochem. Soc. 2015, 162, 436–441. [Google Scholar] [CrossRef]
  49. Aparnev, A.I.; Loginov, A.V.; Uvarov, N.F.; Ponomareva, V.G.; Bagryantseva, I.N.; Manakhov, A.A.; Al-Qasim, A.S.; Golovakhin, V.V.; Bannov, A.G. Novel Highly Dispersed Additive for Proton-Conducting Composites. Appl. Sci. 2023, 13, 5038. [Google Scholar] [CrossRef]
  50. Gusev, A.I. Nanocrystalline Materials: Methods of Preparation and Properties; Ural Branch of the Russian Academy of Sciences: Yekaterinburg, Russia, 1998; 198p. [Google Scholar]
  51. Yong, Z.J. A general sonochemical Approach to Rapid Synthesis of 1D Single-crystalline MSn(OH)6 (M = Ba, Ca, Sr) Nanostructures. Adv. Mater. Res. 2011, 295–297, 1554–1559. [Google Scholar] [CrossRef]
  52. Wang, G.; Bai, J.; Shan, C.; Zhang, D.; Lu, N.; Liu, Q.; Zhou, Z.; Wang, S.; Liu, C. Synthesis and ethanol gas sensing properties of mesoporous perovskite-type BaSnO3 nanoparticles interconnected network. Mater. Lett. 2017, 205, 169–172. [Google Scholar] [CrossRef]
  53. Loginov, A.V.; Aparnev, A.I.; Uvarov, N.F. Synthesis of SrSnO3/SnO2 Composites via Thermal Decomposition of a Precursor. Inorg. Mater. 2022, 58, 420–424. [Google Scholar] [CrossRef]
  54. Haase, P.; Christensen, H.G.; Nielsen, U.G.; Koch, C.B.; Galazka, Z.; Majzlan, J. Stability and solubility of members of tin perovskites in the schoenfliesite subgroup, □2(BSn4+)(OH,O)6 (B = Ca, Fe3+, Mg, Mn2+, Zn, Cu). Chem. Thermod. Therm. Anal. 2021, 1–2, 100005. [Google Scholar] [CrossRef]
  55. Stanulis, A.; Sakirzanovas, S.; Van Bael, M.; Kareiva, A. Sol–gel (combustion) synthesis and characterization of different alkaline earth metal (Ca, Sr, Ba) stannates. J. Sol-Gel Sci. Technol. 2012, 64, 643–652. [Google Scholar] [CrossRef]
  56. Hu, X.; Lv, G.; Jia, Z.; Jiang, J.; Xiao, T.; Yuan, M.; Tang, Y. A general sonochemical approach to rapid synthesis of 1D single-crystalline MSn(OH)6 (M = Ba, Ca, Sr) nanostructures. Appl. Surf. Sci. 2011, 257, 9008–9013. [Google Scholar] [CrossRef]
  57. Medvedev, A.G.; Mikhaylov, A.A.; Shames, A.I.; Ilyukhin, A.B.; Churakov, A.V.; Grishanov, D.A.; Mel’nik, E.A.; Tripol’skaya, T.A.; Lev, O.; Prikhodchenko, P.V. Identification of barium hydroxo-hydroperoxostannate precursor for low-temperature formation of perovskite barium stannate. Inorg. Chem. 2020, 59, 18358–18365. [Google Scholar] [CrossRef] [PubMed]
  58. Azad, A.-M.; Hashim, M.; Baptist, S.; Badri, A.; Haq, A.U. Phase evolution and microstructural development in sol-gel derived MSnO3 (M = Ca, Sr and Ba). J. Mater. Sci. 2000, 35, 5475–5483. [Google Scholar] [CrossRef]
  59. Loginov, A.V.; Aparnev, A.I.; Uvarov, N.F. Nanocomposites Prepared via Thermal Decomposition of Calcium Hydroxystannate CaSn(OH)6. Inorg. Mater. 2022, 58, 814–821. [Google Scholar] [CrossRef]
  60. Lavrova, G.V.; Shutova, E.S.; Ponomareva, V.G.; Dunyushkina, L.A. Proton conductivity and interphase interaction in CsH2PO4-SrZrO3 composites. Russ. J. Electrochem. 2013, 49, 718–724. [Google Scholar] [CrossRef]
  61. Wang, Y.; Chesnaud, A.; Bevillon, E.; Dezanneau, G. Properties of Y-doped BaSnO3 proton conductors. Solid State Ionics 2012, 214, 45–55. [Google Scholar] [CrossRef]
  62. Prakash, A.; Xu, P.; Faghaninia, A.; Shukla, S.; Ager, J.W.; Lo, C.S.; Jalan, B. Wide bandgap BaSnO3 films with room temperature conductivity exceeding 104 S cm−1. Nat. Commun. 2017, 8, 15167. [Google Scholar] [CrossRef]
  63. Geneste, G.; Dezanneau, G. Competition between elastic and chemical effects in the doping, defect association, and hydration of barium stannate. Solid State Ionics 2017, 308, 121–132. [Google Scholar] [CrossRef]
  64. Buscaglia, M.T.; Leoni, M.; Viviani, M.; Buscaglia, V.; Martinelli, A.; Testino, A.; Nanni, P. Synthesis and characterization of BaSn(OH)6 and BaSnO3 acicular particles. J. Mater. Res. 2003, 18, 560–566. [Google Scholar] [CrossRef]
  65. Wu, X.; Lv, G.; Hu, X.; Tang, Y. A Two-Step Method to Synthesize BaSn(OH)6 Crystalline Nanorods and Their Thermal Decomposition to Barium Stannate. J. Nanomater. 2012, 2012, 912731. [Google Scholar] [CrossRef]
  66. Kashin, A.S.; Ananikov, V.P. A SEM study of nanosized metal films and metal nanoparticles obtained by magnetron sputtering. Russ. Chem. Bull. 2011, 60, 2602–2607. [Google Scholar] [CrossRef]
  67. Kachala, V.V.; Khemchyan, L.L.; Kashin, A.S.; Orlov, N.V.; Grachev, A.A.; Zalesskiy, S.S.; Ananikov, V.P. Target-oriented analysis of gaseous, liquid and solid chemical systems by mass spectrometry, nuclear magnetic resonance spectroscopy and electron microscopy. Russ. Chem. Rev. 2013, 82, 648–685. [Google Scholar] [CrossRef]
  68. Lowell, S.; Shields, J.E.; Thomas, M.A.; Thommes, M. Characterization of Porous Solids and Powders: Surface Area, Pore Size and Density. In Particle Technology Series; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2004; 365p. [Google Scholar] [CrossRef]
  69. Karnaukhov, A.P. Adsorption. In The Texture of Dispersed and Porous Materials; Nauka: Novosibirsk, Russia, 1999; 470p. [Google Scholar]
  70. Cerda, J.; Arbiol, J.; Dezanneau, G.; Diaz, R.; Morante, J.R. Perovskite-Type BaSnO3 Powders for High Temperature Gas Sensor Applications. Sens. Actuators B. 2002, 84, 21–25. [Google Scholar] [CrossRef]
  71. Ahmed, J.; Blakely, C.K.; Bruno, S.R.; Poltavets, V.V. Synthesis of MSnO3 (M = Ba, Sr) nanoparticles by reverse micelle method and particle size distribution analysis by whole powder pattern modeling. Mater. Res. Bull. 2012, 47, 2282–2287. [Google Scholar] [CrossRef]
  72. Lu, W.; Schmidt, H. Lyothermal synthesis of nanocrystalline BaSnO3 powders. Ceram. Inter. 2008, 34, 645–649. [Google Scholar] [CrossRef]
  73. Socrates, G. Infrared and Raman Characteristic Group Frequencies Contents: Tables and Charts; Wiley: Chichester, NY, USA, 2001; 347p. [Google Scholar]
  74. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds; Wiley: New York, NY, USA, 1997; Volume 479. [Google Scholar]
  75. Amalric-Popescu, D.; Bozon-Verduraz, F. Infrared studies on SnO2 and Pd/SnO2. Catal. Today 2001, 70, 139–154. [Google Scholar] [CrossRef]
  76. Shukla, S.S. Synthesis and Dispersion of Barium Stannate Nanopowders. 2011. 28p. Available online: https://core.ac.uk/reader/53187964 (accessed on 1 October 2020).
  77. Baranov, A.I. Crystals with disordered hydrogen-bond networks and superprotonic conductivity. Review. Crystallogr. Rep. 2003, 48, 1012–1037. [Google Scholar] [CrossRef]
Figure 1. Synthesis scheme of the precursor.
Figure 1. Synthesis scheme of the precursor.
Jcs 07 00469 g001
Figure 2. Experimental X-ray diffractogram of the sample (black points) in comparison with the diffractogram calculated using the Rietveld method (red line). The green line is a difference between the experimental data and the calculated curve. Theoretical diffractograms of BaSn(OH)6 and BaCO3 are also presented for comparison.
Figure 2. Experimental X-ray diffractogram of the sample (black points) in comparison with the diffractogram calculated using the Rietveld method (red line). The green line is a difference between the experimental data and the calculated curve. Theoretical diffractograms of BaSn(OH)6 and BaCO3 are also presented for comparison.
Jcs 07 00469 g002
Figure 3. (a) Scanning electron microscopy images of freshly deposited BaSn(OH)6 and (b) after storage in air for one month.
Figure 3. (a) Scanning electron microscopy images of freshly deposited BaSn(OH)6 and (b) after storage in air for one month.
Jcs 07 00469 g003
Figure 4. The change in mass (TG) and thermal effects (DSC) observed when the initial BaSn(OH)6 sample is heated at a rate of 10 K·min−1. The ion current values of the mass spectrometric sensor corresponding to the concentration of released water and CO2 molecules are shown by blue and light green curves, respectively.
Figure 4. The change in mass (TG) and thermal effects (DSC) observed when the initial BaSn(OH)6 sample is heated at a rate of 10 K·min−1. The ion current values of the mass spectrometric sensor corresponding to the concentration of released water and CO2 molecules are shown by blue and light green curves, respectively.
Jcs 07 00469 g004
Figure 5. X-ray diffraction patterns of the initial sample (curve 1) and thermolysis products obtained at temperatures of 500 (curve 2), 600 (curve 3), and 700 °C (curve 4).
Figure 5. X-ray diffraction patterns of the initial sample (curve 1) and thermolysis products obtained at temperatures of 500 (curve 2), 600 (curve 3), and 700 °C (curve 4).
Jcs 07 00469 g005
Figure 6. (a,b) Scanning electron microscopy images of the BaSnO3 sample obtained after heating BaSn(OH)6 at 700 °C at various magnifications.
Figure 6. (a,b) Scanning electron microscopy images of the BaSnO3 sample obtained after heating BaSn(OH)6 at 700 °C at various magnifications.
Jcs 07 00469 g006
Figure 7. FTIR absorption spectra of BaSn(OH)6 (1) and BaSnO3 (2).
Figure 7. FTIR absorption spectra of BaSn(OH)6 (1) and BaSnO3 (2).
Jcs 07 00469 g007
Figure 8. Temperature dependence of the conductivity of pure CsHSO4 and composite systems (1 − x)CsHSO4-xBaSnO3.
Figure 8. Temperature dependence of the conductivity of pure CsHSO4 and composite systems (1 − x)CsHSO4-xBaSnO3.
Jcs 07 00469 g008
Figure 9. X-ray diffraction patterns of composites (1 − x)CsHSO4-xBaSnO3 of different compositions in comparison with the parent compounds.
Figure 9. X-ray diffraction patterns of composites (1 − x)CsHSO4-xBaSnO3 of different compositions in comparison with the parent compounds.
Jcs 07 00469 g009
Table 1. Values of specific surface area and pore size of samples as a function of heating temperature.
Table 1. Values of specific surface area and pore size of samples as a function of heating temperature.
Heating
Temperature, °C
110500600700
Phase compositionBaSn(OH)6, BaCO3,
amorphous phase SnO2 × xH2O
BaCO3,
amorphous phases
BaSnO3 and SnO2
BaSnO3,
BaCO3
(traces)
BaSnO3
Specific surface area,
m2·g−1
59432315
Pore size, nm4.45.83.32.6
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Loginov, A.V.; Aparnev, A.I.; Uvarov, N.F.; Ponomareva, V.G.; Bannov, A.G. Synthesis of BaSnO3 as a Highly Dispersed Additive for the Preparation of Proton-Conducting Composites. J. Compos. Sci. 2023, 7, 469. https://doi.org/10.3390/jcs7110469

AMA Style

Loginov AV, Aparnev AI, Uvarov NF, Ponomareva VG, Bannov AG. Synthesis of BaSnO3 as a Highly Dispersed Additive for the Preparation of Proton-Conducting Composites. Journal of Composites Science. 2023; 7(11):469. https://doi.org/10.3390/jcs7110469

Chicago/Turabian Style

Loginov, Anton V., Alexander I. Aparnev, Nikolai F. Uvarov, Valentina G. Ponomareva, and Alexander G. Bannov. 2023. "Synthesis of BaSnO3 as a Highly Dispersed Additive for the Preparation of Proton-Conducting Composites" Journal of Composites Science 7, no. 11: 469. https://doi.org/10.3390/jcs7110469

Article Metrics

Back to TopTop