Akira
Yoko
*a,
Makoto
Akizuki
a,
Naoto
Umezawa
b,
Takahisa
Ohno
b and
Yoshito
Oshima
a
aDepartment of Environment Systems, Graduate School of Frontier Sciences, The University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa, Chiba 277-8563, Japan. E-mail: yoko@oshimalab.k.u-tokyo.ac.jp; Fax: +81 4 7136 4694; Tel: +81 4 7136 4694
bInternational Center for Materials Nanoarchitectonics (MANA), National Institute for Materials Science, 1-1 Namiki, Tsukuba, Ibaraki 305-0044, Japan
First published on 11th July 2016
We studied the formation of Ba1−xSrxZrO3 (0 ≤ x ≤ 1) nanoparticles under highly super-saturated conditions, using supercritical water. It is known that B-site Zr in the perovskite structure plays a dominant role at the nucleation stage, with high nucleation rates under supercritical conditions; this is due to the significantly lower solubility of Zr, compared with A-site ions (i.e., Zr precipitates faster, and A-site ions are taken up into particle after the Zr nucleation). However, in this study, it was found that A-site Ba and Sr significantly influenced the particle size, the A-site deficiency rate, and the surface-OH density of the nanoparticles. The differences in particle size suggested that the ripening or coalescence that occurred after the nucleation stage was dominant in determining the particle size, even under highly super-saturated conditions such as those realized in the supercritical hydrothermal synthesis. The characteristic nanostructure formed in the supercritical water was analyzed in detail; variables such as the A-site deficiency rate and the surface-OH density were investigated. The existence of vacancies at the A-site was confirmed using X-ray absorption fine structure, and a highly defective structure was obtained, particularly when the Ba content was high. The surface state of the nanoparticles was also studied using X-ray photoelectron spectroscopy and first-principles calculations, with the aim of understanding the differences in particle size, and the effects of the A-site deficiencies; the amount of surface-OH corresponded to the A-site deficiency rate, and had an inverse relationship with the particle size.
Supercritical hydrothermal synthesis is a solution method that can be applied for the synthesis of metal oxide nanoparticles with a narrow size distribution, and good crystallinity. Nanoparticles can be obtained using supercritical water, because of the acceleration of nucleation and the inhibition of ripening. High nucleation rates can be realized by using a high degree of super-saturation, and ripening can be suppressed via an extremely low solubility in supercritical water.7 A high degree of super-saturation can be obtained when the temperature of the water is increased rapidly from a normal temperature to a temperature above the critical temperature; the low kinetic viscosity of supercritical water enables rapid mixing of fluids with a continuous flow reactor that leads to a rapid increase in water temperature. Higher-crystallinity particles can be obtained, compared with other solution methods, because of the higher reaction temperature. The high crystallinity of the synthesized particles means that the calcination post-processing that is required for conventional solution methods is not needed, or can be simplified, when supercritical hydrothermal methods are applied to perform granulation.8
A distinctive feature of supercritical hydrothermal synthesis is the rapid crystallization that occurs as a result of the high degree of super-saturation. Semi-stable-state or non-equilibrium-state materials can be formed during such rapid structural formation; highly super-saturated conditions can affect the structural and compositional factors in addition to the particle size. For instance, we previously studied the supercritical hydrothermal synthesis of BaZrO3, and discovered the formation of BaZrO3, which had an extremely high concentration of deficiencies at the A-site in the first stage of crystallization.9,10 The formation of nanoparticles in supercritical water with a similarly deficient structure was also reported in a previous study on the mechanism of formation of nickel ferrite.11 The non-stoichiometry and defective structures observed in these studies can affect the properties of particles, and the structural analysis of nanoparticles formed in supercritical water is important. Particularly in the case of composite oxides, the non-stoichiometry and the deficient local structure becomes more crucial during the synthesis, because each element has a different driving force of crystallization, because of the differences in their solubility in water.
In this study, the supercritical hydrothermal synthesis of Ba1−xSrxZrO3 (0 ≤ x ≤ 1) was carried out, and structural analysis of the products was performed to elucidate the formation of the structure in the highly supersaturated reaction field. The perovskite-type oxides BaZrO3 and SrZrO3 are expected to find applications as proton conductors.12–14 Composite oxides containing both Ba and Sr at the A-site show dielectric properties,15–17 because the crystal structure of zirconate varies depending on its composition; some structural studies have also been conducted investigating the phase transition18,19 and the thermal properties.20
We previously investigated the mechanism of formation of BaZrO3 in supercritical hydrothermal synthesis, and observed the existence of a highly Ba-deficient perovskite structure, which is a characteristic of the methodology.9,10 These results suggested that extraordinary structures could appear in supercritical water, due to the accelerated crystallization. Previous studies have shown that A-site-deficient perovskite oxide is important for a variety of applications. Several studies have been performed, including investigations of strontium niobium cobalt oxide (Sr0.95Nb0.1Co0.9O3−δ) for low-temperature solid-oxide fuel cells,21 the deficient effects of yttrium-doped barium cerate zirconate (BaxCe0.5Zr0.4Y0.1O3−δ (x = 0.95–1.05)) on sinterability and electrical conductivity,22 gadolinium tantalite (Gd1/3TaO3) for lithium insertion reactions,23,24 and barium lanthanum titanate (Ba(1−x)/2 Lax/3TiO3 (x = 0.1–1.0)) for dielectric materials.25 A detailed analysis of the products using X-ray diffraction (XRD), transmission electron microscopy energy-dispersive X-ray spectroscopy (TEM-EDX), X-ray photoelectron spectroscopy (XPS), X-ray absorption fine structure (XAFS), and first-principles calculations based on density functional theory (DFT), was conducted to investigate the mechanism of formation of Ba1−xSrxZrO3 (0 ≤ x ≤ 1) in supercritical water.
Ba starting material | Sr starting material | Zr starting material | Base solution | x | |||||
---|---|---|---|---|---|---|---|---|---|
Ba(NO3)2 | 0.020 M | Sr(NO3)2 | — | ZrO(NO3)2 | 0.005 M | KOH | 0.05 M | 4 | 0 |
Ba(NO3)2 | 0.015 M | Sr(NO3)2 | 0.005 M | ZrO(NO3)2 | 0.005 M | KOH | 0.05 M | 4 | 0.25 |
Ba(NO3)2 | 0.010 M | Sr(NO3)2 | 0.010 M | ZrO(NO3)2 | 0.005 M | KOH | 0.05 M | 4 | 0.50 |
Ba(NO3)2 | 0.009 M | Sr(NO3)2 | 0.011 M | ZrO(NO3)2 | 0.005 M | KOH | 0.05 M | 4 | 0.55 |
Ba(NO3)2 | 0.008 M | Sr(NO3)2 | 0.012 M | ZrO(NO3)2 | 0.005 M | KOH | 0.05 M | 4 | 0.60 |
Ba(NO3)2 | 0.007 M | Sr(NO3)2 | 0.013 M | ZrO(NO3)2 | 0.005 M | KOH | 0.05 M | 4 | 0.65 |
Ba(NO3)2 | 0.005 M | Sr(NO3)2 | 0.015 M | ZrO(NO3)2 | 0.005 M | KOH | 0.05 M | 4 | 0.75 |
Ba(NO3)2 | — | Sr(NO3)2 | 0.020 M | ZrO(NO3)2 | 0.005 M | KOH | 0.05 M | 4 | 1 |
The solid products were recovered via pressurized filtering using a nitrocellulose filter (VSWP14250; Merck Millipore Corporation, Darmstadt, Germany, pore size: 0.025 μm (we confirmed that single nanometer particles are also recovered by this filter)9), and were dried in a vacuum oven at room temperature (ADP-31; Yamato Corporation, Tokyo, Japan). The recovered particles and filtrate were analyzed as described below.
![]() | ||
Fig. 2 Relationship between the x value of products and the feed solutions (x in Ba1−xSrxZrO3), measured using ICP-AES for filtrates (rhombuses), and TEM-EDX for each single nanoparticle (circles). |
Fig. 3 shows the molar ratio of A-site ions per B-site ion, obtained using TEM-EDX. Plots show the averaged value of 20 measurements, and error bars show standard deviation of the measured values which were approximately 10–25% of (Ba + Sr)/Zr ratio. The existence of more A-site vacancies in BaZrO3 (x = 0) was suggested, and fewer A-site vacancies were expected when the amount of Sr at the A-sites was increased. Basically, Sr is easier to incorporate into the structure as shown in Fig. 3 due to a smaller ionic radius of Sr than that of Ba. However, Fig. 2 shows that Ba is more incorporated specifically at x = 0.25, and we think that structure difference affected the incorporation (i.e. Ba is more incorporated into cubic perovskite, and Sr is more incorporated into orthorombic perovskite). This tendency was confirmed using ICP-AES for the filtrates and XPS for the nanoparticles, as shown in Fig. 4. It should be noted that the values obtained using XPS were always smaller than the values obtained using TEM-EDX; this outcome suggested the existence of a Zr-rich phase (i.e., an A-site-defective phase), in the surface of the nanoparticles.
![]() | ||
Fig. 3 Molar ratio of A-site ions (Ba + Sr) per B-site ion (Zr), obtained using TEM-EDX (rhombuses: (Ba + Sr)/Zr, up-triangles: Sr/Zr, down-triangles: Ba/Zr). |
![]() | ||
Fig. 4 Molar ratio of A-site ions (Ba + Sr) per B-site ion (Zr), obtained using ICP (triangles), XPS (circles), and TEM-EDX (squares). |
![]() | ||
Fig. 5 Rietveld fitting results around the structural phase transition regions: (A) x = 0.25 (Pm![]() |
Fig. 6 illustrates the dependence of the lattice volume on the A-site composition; the lattice volume decreased with increases in x. The ionic radius of Ba2+ (142 pm) is larger than that of Sr2+ (126 pm), and the changes in the lattice parameter could be explained by the differences in the ionic radii. Fig. 6 also shows the lattice volume obtained via first-principles calculation. A decrease in lattice volume with increases in Sr was observed in both the experimental and the computational data, suggesting that both Ba and Sr were located at the A-sites of the perovskite, and that the presence of A-site ions significantly affected the lattice volume. A slight divergence from Vegard's law was observed in the experimental results, while the computational results with full relaxation clearly obeyed Vegard's law. The variance in lattice volume could have been caused by the difference in particle size; this is discussed in Section 3.4.
![]() | ||
Fig. 6 Lattice volume for each x value, compared with values obtained using the DFT-PBE calculations, (squares) and values refined by Rietveld analysis (triangles). |
![]() | ||
Fig. 7 (A) k3χ(k) spectrum obtained from Zr K edge absorption spectrum; (B) radial structure function obtained via a Fourier transform of the k3χ(k) spectrum. |
![]() | ||
Fig. 8 Coordination numbers of Ba that second coordinated to Zr, obtained from the fitting for the radial structure function. |
β2(cos![]() ![]() |
![]() | ||
Fig. 9 (A) Crystallite size at each x value of Ba1−xSrxZrO3, obtained using the Williamson–Hall method. (B) Particle size of Ba1−xSrxZrO3 observed using TEM. |
![]() | ||
Fig. 10 (A) TEM images for x = 0 and x = 1; (B) coefficient of variation of particle size observed by TEM. |
The composition of the A-sites influenced the particle size, although the B-site Zr has a dominant role at nucleation stage (i.e., Zr precipitated faster than the A-site ions, because of the much lower solubility of the Zr, and A-site ions were taken up into the particles after the Zr nucleation). It was therefore suggested that the particle size was determined not only by the nucleation of Zr, but also by the subsequent stage involving the formation of the composite oxide phase.
It should be noted that, here, “subsequent stage” does not refer to Ostwald ripening. When a particle grows via Ostwald ripening, the particle size distribution tends to become narrower with increases in particle size; however, such a tendency was not observed in our TEM observations, as demonstrated by the x-dependence of the coefficient of variation for the particle size in Fig. 10B. Considering that each particle was a mono-crystal, it was suggested that the growth could have been related to the coalescence of nuclei, which occurs only at the early stage of crystallization, when each particle has enough deformable structure to grow as a result of collisions with other particles.
Philippot et al. synthesized Ba1−xSrxTiO3 in supercritical water, and suggested that the surface state, particularly the hydroxyl groups (–OH) on the particle surface, had some influence on the increase in particle size. They concluded that the concentration of surface-OH can be regarded as an index for ripening.33 Here, XPS analysis was conducted to investigate whether the difference in particle size could be attributed to the surface density of –OH in this case. First, the peak positions in the O 1s spectra were calculated, and peak-fitting analysis was then conducted to elucidate the differences in surface state.
First-principles calculations were conducted to confirm the relationship between crystal structure and the shift in binding energy for each sample. The core level shift of O 1s was obtained using an initial state approximation implemented in VASP.34 The Kohn–Sham energy of the core states was obtained, and the core level shift was calculated referring to the Fermi level, as shown in the following equation:
ΔE(core level shift) = (εO1s(x = 0.25, 0.50, 0.75, 0.875, and 1) − εF(x = 0.25, 0.50, 0.75, 0.875, and 1)) − (εO1s(x = 0) − εF(x = 0)) | (1) |
Fig. 11 shows the experimental binding energy difference and simulated core level shift in O 1s for each x in Ba1−xSrxZrO3. From the results, it was confirmed that structural differences resulting from changes in x caused the shifts in the binding energy. The abrupt drop in the core level shift around x = 0.75 observed in the experimental results might have originated from the local atomic displacement. Fig. 11 also shows computational results achieved using atomic positions determined using diffraction data (denoted as unrelaxed structure in the following); a similar inverse shift was observed at x = 0.75. To analyze this phenomenon, the distortion index proposed by Baur35 for the ZrO6 polyhedron was estimated for the full relaxed structure, 0.0047, and for the unrelaxed structure, 0.0366, at x = 0.75. The distortion index represented the variance of the bond length, and the larger the index, the more distortion remained in the structure, indicating that the unrelaxed structure held a larger amount of distortion. Because the abrupt drop in the core level shift at x = 0.75 was observed in the computational results with unrelaxed structure, and in the experimental results, we conclude that the products at x = 0.75 were subject to larger distortion in the crystal, and that the local distortion led to the irregular shift in binding energy. In addition, it was confirmed that the effects of the A-site configuration on the core level shift were negligible, from the comparison of the results given by two different configuration models (Model 1 and Model 2). These simulations provided a validation for the shift in peak positions in our peak separation fitting for the surface-OH and structural-O.
![]() | ||
Fig. 11 Core level shift of O 1s for each x for Ba1−xSrxZrO3 obtained by XPS (squares), DFT-PBE without relaxation (circles), and DFT-PBE with full relaxation (triangles). |
Fig. 12 shows the O 1s peaks of the XPS spectra for each x, after subtraction of the background, with three fitting curves for bulk-O (528.5–529.2 eV), surface-OH (530.3–531.1 eV), and absorbed water (531.6–533.0 eV). Multi-peak fitting was conducted, to divide peaks into the three components using the Voigt function.
Fig. 13 shows the peak area ratio of surface-OH to structural-O at each x value, obtained via the fitting. In contrast with the previous study by Philippot et al.,33 the density of surface-OH decreased with increases in the molar ratio of Sr, i.e., the increase in particle size. In this case, the difference in particle size could not be explained by the higher density of surface-OH; rather, it is likely that collision growth was prevented by the larger repulsive forces resulting from the high density of surface-OH. A higher concentration of A-site deficiencies and a higher density of surface-OH was observed under Ba-rich conditions; it is believed that the charge neutrality for the A-site deficiencies was compensated for by the inclusion of protons in the crystal, as mentioned in Section 3.3. Further study of the structural formation, considering in particular the surface structure of the nanoparticles, is required to elucidate the growth mechanism and utilize these nanoparticles as new materials.
![]() | ||
Fig. 13 Intensity ratio of surface-OH per structural-O, obtained using the fitting results at each x value. |
The formation of nanoparticles of Ba1−xSrxZrO3 (0 ≤ x ≤ 1) during supercritical hydrothermal synthesis was investigated for the first time. Composite oxides including both Ba and Sr at the A-sites were obtained, and it was found that the A-site composition greatly affected the crystal structure, particle size, vacancy concentration, and surface state. A-site ions played a particularly important role in determining the particle size. High concentrations of A-site deficiencies and a corresponding increase in surface-OH in Ba-rich products were related to the prevention of particle growth. In addition, it is plausible that the diffusibility of A-site ions in the solid determined the concentration of A-site vacancies, and that protons compensated for the charge of the A-site deficiencies, resulting in increases in the concentration of surface-OH with the increases in the number of A-site deficiencies.
Regarding the supercritical hydrothermal synthesis of zirconate, zirconium precipitation is so rapid that all zirconium ions turn into a solid phase at an early stage of crystallization, due to the extremely low solubility of zirconium in supercritical water. Following the Zr nucleation, the crystallization in the particles proceeds, and a composite oxide phase is formed via the uptake of Ba and Sr. In this model, Zr, which has a lower solubility, takes a leading role in the nucleation and growth. However, in this study it was found that the zirconate particle size differed depending on the A-site composition. The difference in the particle size could not be explained sufficiently by considering only Zr nucleation and A-site ion (Ba, Sr) uptake. This indicated that a subsequent process of particle growth following the nucleation process, such as ripening or coalescence, was a key phenomenon in controlling the particle size. The investigation of the particle size variance indicated that a coalescence mechanism occurring only at the early stages of crystallization was plausible. Abundant A-site deficiencies and surface-OH could have prevented collision particle growth. It was concluded that the coalescence mode in the early stages of crystallization affected the products, even under highly super-saturated conditions in supercritical water, which accelerated nucleation and inhibited growth.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra12288d |
This journal is © The Royal Society of Chemistry 2016 |