Alejandro Gómez-Péreza,
Jesús Prado-Gonjalb,
Daniel Muñoz-Gilb,
Adrián Andrada-Chacónc,
Javier Sánchez-Benítezc,
Emilio Moránb,
María Teresa Azcondoa,
Ulises Amadora and
Rainer Schmidt*d
aDpto. Química, Facultad de Farmacia, Universidad San Pablo – CEU, 28668 Boadilla del Monte, Madrid, Spain
bDpto. Química Inorgánica, Facultad de CC. Químicas, Universidad Complutense de Madrid, 28040 Madrid, Spain
cDpto. Química Física I, Facultad de CC. Químicas, Universidad Complutense de Madrid, 28040 Madrid, Spain
dDpto. Física Aplicada III, Facultad de CC. Físicas, GFMC, Universidad Complutense de Madrid, 28040 Madrid, Spain. E-mail: rainerxschmidt@googlemail.com
First published on 2nd October 2015
In this work we report on the microwave assisted synthesis of nano-sized Gd2Ti2O7 (GTO) and Ho2Ti2O7 (HTO) powders from the RE2Ti2O7 pyrochlore family (RE = rare earth). Synchrotron X-ray powder diffraction was used to study RE–Ti cationic anti-site defects with concentrations that decrease in both samples with increasing temperature starting from 1100 °C, and the defects disappear at 1400 °C. SQUID magnetometry measurements revealed that GTO shows a predominantly anti-ferromagnetic structure, whereas HTO exhibits magnetic saturation and a ferromagnetic component at low temperature. Impedance spectroscopy data revealed strongly increased ionic oxygen vacancy conduction in HTO ceramic pellets as compared to GTO, which may be associated with a higher degree of oxygen vacancy disorder. This argument was supported by Raman spectroscopy data.
Rare-Earth (RE) titanate pyrochlores have the general formula RE2Ti2O7, where the crystal structure may be regarded a RE3+ and Ti4+ cation ordered superstructure of oxygen deficient fluorite (RE0.5Ti0.5O2-0.25). A concentration of 0.25 oxygen vacancies per formula unit allows ionic oxygen vacancy conduction to occur. The pyrochlore structure (Fig. 1) exhibits an F-centred cubic unit cell with space group (S.G): Fdm (#227) and cell parameter a ≈ 10 Å, which corresponds to approximately the double of the fluorite unit cell parameter.10 The larger RE3+ cations are located on 16d eight-coordinated sites at the centre of scalenohedrons, while Ti4+ are located on 16c six-coordinated sites at the centre of trigonal antiprisms. The 7 oxygen O2− anions in the chemical formula RE2Ti2O6O′ are located on two different sites: (1) 6 oxygen O(1) on four-coordinated 48f positions with two A cation and two B cation nearest neighbours, and (2) one oxygen O(2) on four-coordinated 8b sites with four A cation nearest neighbours. The third anionic tetrahedral O(3) sites are empty, consisting in four-coordinated 8a sites surrounded by four B cation neighbours. These systematically vacant O(3) sites are characteristic of the ordered pyrochlore structure, in contrast to anion deficient fluorites where anion vacancies are randomly distributed throughout the anion sub-structure. Merely based on steric considerations regarding the cationic radii r, one would expect the eight-coordinated 16d sites in RE2Ti2O7 pyrochlores to be fully occupied by the larger RE3+ cations, whereas the smaller Ti4+ would be located on the six-coordinated 16c positions: r(VIIIGd3+) = 1.053 Å, r(VIIIHo3+) = 1.015 Å and r(VITi4+) = 0.605 Å.11 However, in the A2B2O7 pyrochlore structure crystal defects commonly occur, where the two most important ones are (i) cation anti-site, and (ii) anion Frenkel defects. Cation anti-site defects consist in a positional exchange of an A3+ (here RE3+) with a B4+ (here Ti4+) cation, whereas Frenkel anion disorder occurs when a 48f site oxygen anion is displaced to a nominally vacant 8a interstitial site, thus leaving a 48f site unoccupied.10,12
Disordered oxygen vacancies (e.g. in oxygen deficient fluorites) favour higher oxygen ion mobility and conductivity.13–16 Therefore, high oxygen mobility and conductivity in the pyrochlore structure was claimed to be favoured by a certain degree of A and B anti-site defects, because anti-site defects may be directly coupled to the oxygen vacancy disorder, i.e. increased anti-site defects directly induce increased oxygen vacancy disorder in the nominally oxygen ordered structure and improve oxygen ion conductivity.13–15 In this context it should be noted that increasing concentrations of pyrochlore A and B anti-site defects can be regarded a precursor to the formation of the fluorite structure with a disordered arrangement of the cationic sub-lattice and the oxygen vacancies. On the other hand, some authors have reported that some degree of oxygen structural disorder can exists in pyrochlores despite a fully ordered cationic sub-lattice, as evidenced by Raman spectroscopy, and the anti-site and anion Frenkel defects may not always be directly coupled.17,18
Here in this work the coupling of anti-site and Frenkel defects could not be confirmed for the HTO samples, because a negligible anti-site defect concentration was detected by synchrotron X-ray diffraction (SXRD), but Raman spectroscopy suggested considerable oxygen vacancy disorder in HTO. The latter finding was corroborated by significantly enhanced ionic oxygen vacancy conduction in the HTO sample as compared to GTO, detected by impedance spectroscopy on sintered ceramic pellets.
It has been established previously that the concentration of pyrochlore A and B anti-site defects depends on the details of the synthesis procedure.17–19 Here in this work we demonstrate that the microwave (MW)-assisted method produces off-equilibrium (metastable) pyrochlore-type phases with varying amounts of anti-site defects, in agreement with other more recently developed alternative synthesis methods, e.g. mechanical milling.19 Traditionally, RE2Ti2O7 compounds are prepared by solid state reactions, which require repeated and long synthesis procedures at temperatures of up to ≈1400 °C and intensive grinding at the initial and intermediate stages of processing to achieve single phase material.20–22 In the last decades several alternative routes have been proposed such as microwave (MW) synthesis, molten salt mediated synthesis, co-precipitation of hydroxides, sol–gel and citrate based methods, and mechanical milling, where the time and temperature requirements for synthesis are significantly reduced.19,23–26 However, these methods usually require a second calcination step in a conventional furnace to obtain phase pure materials and the total reaction time increases. MW-assisted synthesis may be particularly useful due to the nanometre-sized particles that can be obtained at reduced synthesis time and processing costs.27–32 In the work presented here we use a domestic microwave oven for the first heat treatment process of only 20 min, followed by a 2 h calcination procedure at 1100 °C in a conventional furnace to obtain highly crystalline single phase materials.33 We explore this fast MW-assisted synthetic route for RE2Ti2O7 pyrochlores with RE = Gd, Ho. The nano-metric particle sizes may improve the sintering activity of the resulting powders, which is beneficial to obtain dense bulk materials by ceramic sintering to study the dielectric properties. Furthermore, a quick and economic sintering process may be relevant for potential industrial applications. By example, we present the microwave synthesis of Gd2Ti2O7 (GTO) and Ho2Ti2O7 (HTO) pyrochlores, and their structural, magnetic and dielectric properties.
In the field of solid state synthesis especially domestic microwave furnaces are well-known to often entail problems with the reproducibility of the final products due to the lack of temperature control. In this work we did not encounter this problem, most likely due to the fact that we used a microwave heating source only for the first synthesis step to produce precursor powder followed by a second step calcination process in a conventional furnace. In the conventional furnace the temperature is well-controlled and reproducibility problems usually do not occur.
All samples were analysed by standard XRD and scanning electron microscopy (SEM) twice: first, after the initial microwave treatment and second, after the subsequent calcination treatment at 1100 °C in the conventional furnace. In each case the pellets were extracted from the microwave or the conventional furnaces and were ground into powder form. Three batches of the final single phase products were further heated at 1200 °C, 1300 °C and 1400 °C for 12 h each to study the temperature dependence of the concentration of anti-site defects. For ionic conductivity measurements by impedance spectroscopy, GTO and HTO pressed pellets of 5 mm diameter and 2 mm thickness were sintered at 1400 °C for 4 h (heating and cooling ramps of 2 °C min−1). The relative densities of the sintered GTO and HTO pellets were ≈95% of the crystallographic value. Both faces of the pellets were covered with Au electrodes using dc sputtering to facilitate impedance spectroscopy measurements.
High-resolution powder synchrotron X-ray diffraction (SXRD) patterns were collected at the SpLine beamline (BM25A) of the Spanish CRG at the European Synchrotron Radiation Facility (ESRF, Grenoble) with a fixed wavelength of λ = 0.49684897 Å (24954.112 eV). Powdered samples were placed inside capillaries with diameters ranging from 0.2–0.5 mm, where the capillaries were rotated during exposure. In each case the absorption was calculated to maximize the signal-to-time ratio for every sample. Data were collected in a continuous 2θ-scan mode from 3° to 35°.
Structural refinements using (S)XRD data were carried out by the Rietveld method using FULLPROF software.34 A Pseudo-Voigt function was selected as the line profile and the corresponding parameters were refined. Furthermore, the set of pre-defined background points, scale factor, zero shift, lattice parameters, atomic O displacement and the overall Debye–Weller factors were refined. These overall factors were converted into individual isotropic Debye–Waller factors and refined. The cation occupation was also refined to detect possible anti-site defects.
Scanning electron microscopy (SEM) micrographs were collected on an FEI XL30® apparatus equipped with an attached analyser for energy dispersive spectroscopy (EDS).
Unpolarised Raman spectroscopy experiments were performed using a micro-Raman confocal spectrometer (Voyage™ BWS435-532SY, BW&Tek). An excitation wavelength of 532.0 nm was produced by a solid state laser with an incident power fixed at 10 mW. The spectrometer was equipped with an Olympus BX51 microscope allowing the collection of the Raman signal on backscattering configuration through a 50× long working distance objective. With this setup, the spectral resolution was about 4 cm−1. A Chameleon™ CMLN-13S2C digital camera has also been coupled to the equipment.
Magnetic susceptibility measurements of GTO and HTO powder samples calcined at 1100 °C were performed in a Quantum Design XL-SQUID magnetometer in the temperature range of 2–300 K at 1 kOe applied magnetic field. The temperature dependence of the magnetic susceptibility was measured following zero-field-cooled and field-cooled (ZFC–FC) procedures with intermediate demagnetization at room temperature. Ferri-magnetic hysteresis cycles were recorded at 1.7 K from magnetization vs. applied field measurements.
Impedance spectroscopy was carried out on 1400 °C sintered ceramic pellets between 150 and 550 K using a Novocontrol Alpha-A High Performance Frequency Analyzer equipped with a nitrogen gas cooled sample chamber. A 100 mV amplitude alternating voltage signal was employed and the data was recorded between 0.5 Hz to 3 MHz in terms of the real and imaginary parts of the impedance (Z′–Z′′). The data were converted into the permittivity notation ε′–ε′′ using the standard conversion.35 High temperature impedance spectroscopy was carried out using a Solartron 1260 Frequency Response Analyzer in the 1 Hz to 1 MHz range between 298 and 1273 K in different atmospheres (air, pure oxygen or pure Ar), using a 100 mV amplitude alternating voltage signal.
![]() | ||
Fig. 4 SEM micrographs of the RE2Ti2O7 samples. (a) and (b) correspond to Gd2Ti2O7 and Ho2Ti2O7 powder samples. (c) and (d) correspond to the respective pellets sintered at 1400 °C during 4 h. |
The structure factors (Fhkl) of the superstructure peaks depend on the difference of scattering power between the 16d and the 16c sites and the Fhkl factors for peaks with even h, k and l depend on the total of the scattering power in both positions. Therefore, the relative intensities of the superstructure peaks are a direct reflection of the cation distribution in the structure.
We started our structural refinements of the SXRD data with the fully ordered pyrochlore structure, where all RE ions are constrained to the 16d sites and the Ti to the 16c sites. However, it became evident from the fitting process that more scattering power should be located in the 16c site which could only be accomplished by allowing the exchange of RE and Ti cations from their nominal positions, i.e. the occurrence of A and B anti-site defects. The final refined cation distributions presented in Table 1 show a perceptible fraction of Gd and Ho cations to be located at the six-coordinated 16c site, which reduced after heating the powders at higher temperatures (up to 1400 °C). The anti-site defect percentages are 3.6% for Gd2Ti2O7 and 2.4% for Ho2Ti2O7 at 1100 °C. This is consistent with the different relative intensities of the superstructure peaks shown in Fig. 7, where the anti-site defect concentrations show clear temperature dependence. Generally, anti-site defects are more common when the A and B cations have similar ionic radii, whereas in the RE2Ti2O7 case presented here the A- and B-site cationic radii are quite different and the anti-site defect concentration is then expected to be strongly temperature dependent.18,19,36 The temperature dependence of the anti-site defect concentrations is summarized in Table 1.
Ho2Ti2O7 | Gd2Ti2O7 | ||||||
---|---|---|---|---|---|---|---|
1100 °C | 1200 °C | 1400 °C | 1100 °C | 1200 °C | 1300 °C | 1400 °C | |
a (Å) | 10.10004(6) | 10.09882(5) | 10.09828(3) | 10.18126(8) | 10.18407(5) | 10.18496(5) | 10.18507(6) |
A position | 16d | 16d | |||||
U × 100 (Å2) | 0.72(1) | 0.74(2) | 0.38(3) | 0.84(1) | 0.62(2) | 0.56(3) | 0.45(2) |
Occ RE/Ti | 1.952(3)/0.048(3) | 1.952(2)/0.048(2) | 2.000(2)/0.000(2) | 1.929(3)/0.071(3) | 1.952(2)/0.048(2) | 1.976(2)/0.024(2) | 2.000(2)/0.000(2) |
B position | 16c | 16c | |||||
U × 100 (Å2) | 1.01(5) | 1.03(3) | 0.39(2) | 1.22(2) | 0.83(1) | 0.65(2) | 0.95(4) |
Occ Ti/RE | 1.952(3)/0.048(3) | 1.952(2)/0.048(2) | 2.000(2)/0.000(2) | 1.929(3)/0.071(3) | 1.952(2)/0.048(2) | 1.976(2)/0.024(2) | 2.000(2)/0.000(2) |
O(1) position | 48f | 48f | |||||
X | 0.4229(4) | 0.4232(4) | 0.4227(4) | 0.4239(4) | 0.4244(3) | 0.4253(3) | 0.42877(5) |
U × 100 (Å2) | 0.71(2) | 0.71(2) | 0.42(1) | 0.82(3) | 1.13(1) | 0.88(2) | 0.93(1) |
O(2) position | 8b | 8b | |||||
U × 100 (Å2) | 0.09(2) | 0.02(4) | 0.51(3) | 0.09(2) | 0.16(3) | 0.01(2) | 0.23(3) |
χ2 | 4.24 | 4.43 | 4.60 | 4.78 | 4.74 | 3.57 | 6.21 |
Rwp/Rexp (%/%) | 9.63/4.67 | 9.13/4.34 | 9.36/4.36 | 10.7/4.92 | 9.23/4.24 | 7.59/4.02 | 12.3/3.94 |
RBragg | 2.42 | 2.35 | 2.83 | 2.40 | 2.16 | 1.90 | 5.50 |
Space group | Fd![]() |
Fd![]() |
The pyrochlore unit cell parameter is mainly determined by the size of the BO6 distorted octahedra (=trigonal antiprisms) and, bearing in mind the ionic radii of RE3+ and Ti4+ cations, it is clear that a higher degree of occupation of the 16c sites by RE cations increases the size of the BO6 units and concomitantly the unit cell.19 Since thermal treatment removes anti-site defects it would therefore be expected that heating at higher temperatures would result in smaller unit cell parameters. Additionally, the unit cell parameters may also be affected by a small concentration (δ) of non-stoichiometric oxygen vacancies (RE2Ti2O7−δ), compensated by the formation of Ti3+.37,38 By heating at high temperature in oxidizing conditions these anion vacancies would be filled, accompanied by oxidation of large Ti3+ cations to smaller Ti4+, and the unit cell parameter would again decrease. We believe that non-stoichiometric oxygen vacancies may well exist in our samples since we use strongly reducing carbon as MW-absorber and the obtained products would therefore be expected to be oxygen-deficient to a certain extent. Therefore, we presume that at least two mechanisms may operate at the same time during heat treatment at elevated temperatures in air: (i) cation re-distribution among A- and B-sites, and (ii) oxidation of Ti3+ to Ti4+ through incorporation of oxygen into non-stoichiometric anion vacancies. Both processes would induce shortening of the unit cell.
The unit cell size for HTO (Table 1) displays such a trend, whereas for GTO the unit cell shows an expansion as the temperature of the heat treatments increases. We suspect that the thermal expansion of the GTO lattice may dominate over the decreasing anti-site defect concentration and/or the concentration of non-stoichiometric oxygen vacancies may be significantly lower, which then results in a different evolution of the unit cell size. Alternatively, the GTO phase may be affected by additional structural variations.
![]() | ||
Fig. 8 Raman shift. (a) Gd2Ti2O7 (red) heat treated at 1100 °C, (b) Gd2Ti2O7 heat treated at 1400 °C, (c) Ho2Ti2O7 (blue) heat treated at 1200 °C, (d) Ho2Ti2O7 heat treated at 1400 °C. |
The main feature of the Raman spectra shown in Fig. 8 is the appearance of an additional phonon (large blue arrows) centred at 750 cm−1 in HTO (but not in GTO). This mode has been observed previously in Y2Zr2−yTiyO7 pyrochlore structures and was associated with a substructure with varying oxygen coordination.42 We interpret the occurrence of this mode as a clear sign of oxygen disorder in the HTO, and the two types of defects, anti-site and Frenkel anion defects, may be uncoupled to a certain extent. This notion is corroborated by a strongly increased ionic oxygen vacancy conduction in the HTO sample as demonstrated below (Section 3.6) despite the absence of anti-site disorder detected by SXRD measurements (Section 3.3).
Fig. 9 shows the temperature dependence of the magnetic susceptibility for both powder samples. No significant differences between ZFC (solid marks) and FC (hollow marks) susceptibility curves are evident. The insets in Fig. 9 show plots of the reciprocal magnetic susceptibility 1/χ vs. T at low temperature, indicating a deviation from the Curie–Weiss law [χ = C/(T − θ)] below ≈10 K in both samples. The Curie–Weiss fits are indicated by black solid lines, which allowed the extraction of the Curie constant C, related to the effective magnetic moment μeff, and the Weiss constant θ. In the case of GTO a linear regression analysis yields μeff ≈ 5.7 μB/Gd3+ and θ ≈ −9.8 K. A negative θ value suggests that antiferromagnetic interactions may be predominant. For HTO, μeff ≈ 7.3 μB/Ho3+, and θ ≈ +0.2 K. The latter positive value suggests that ferromagnetic interactions may be predominant. The experimental magnetic moments are smaller than the theoretical ones: μeff = 7.9 μB for the free ion 8S7/2 ground state of Gd3+ and 10.6 μB for the 5I8 Ho3+ ground state. This discrepancy between theoretical and experimental magnetic moments may be related to the magnetic frustration commonly encountered in pyrochlore structures.
Fig. 10 presents the magnetization vs. applied field plots at 1.7 K for GTO and HTO, where a small hysteresis is observed. In the case of HTO a magnetic saturation is observed, which confirms the positive value of the Weiss constant and the notion that a soft ferromagnetic component drives the magnetic behaviour. A more detailed investigation into the magnetic structure of the GTO and HTO phases in terms of neutron diffraction experiments may be indicated to elucidate the origin of the ferromagnetic component in HTO, but this goes behind the scope of the work presented here.
![]() | ||
Fig. 10 Magnetic hysteresis loop at 1.7 K in the magnetic field range of −50 to 50 kOe for (a) Gd2Ti2O7 (red circles) and (b) Ho2Ti2O7 (blue squares). Insets: magnification of the hysteresis loops. |
At the low frequency end the data points for the HTO sample in the −Z′′ vs. Z′ notation (Fig. 11b) align linearly in a “pike-like” fashion, which is inconsistent with the conventional RC element model. This behavior is commonly interpreted as a blocking effect of the electrode sample interface (IF).35 This constitutes evidence for ionic charge transport where the blocking effect of the interface occurs as a result of the different types of dominating charge carriers in the Au electrodes (electrons) and in the ceramic (oxygen vacancies). Therefore, ionic oxygen vacancy conductivity in the HTO samples is indicated, and electronic contributions may be small. The charge transport across blocking electrode interfaces is usually by diffusion, which was not accounted for in the equivalent circuit model here, because rather complex circuit components such as Warburg elements are needed to describe diffusion processes by an equivalent circuit component.35 For the GTO sample the interface pike contribution is not visible in Fig. 11a, most likely due to insufficient resolution at the low frequency end. The interface contribution may possibly appear at lower frequency. This may be attributed to the higher resistivity of the GTO sample, which indicates a slower charge transport mechanism at the same temperature, and the same features in the impedance spectra would be visible at lower frequency.
The resulting equivalent circuits used for fitting the data from GTO and HTO ceramic pellets are displayed in the insets of Fig. 11a and b and a good fit to the data can be obtained in the high and intermediate frequency range where the bulk contributions are dominant. It should be noted that the semicircles in Fig. 11 are slightly suppressed in a way that the semicircle center seems to be slightly shifted below the Z′ x-axis. This indicates a certain degree of non-ideality of the dielectric relaxation processes, which can be accounted for by connecting a constant phase element (CPE) in parallel to the ideal RC element.47 The CPE behavior is usually explained in the framework of a broadening of the distribution of relaxation times τ across the macroscopic sample, where τ = RC, with R being the resistance and C the capacitance of an ideal RC element.35,48
In both cases, for GTO and HTO, only one semicircle is evident from Fig. 11, although the pike-like interface contribution suggests ionic type charge transport, at least in the HTO. This is rather unusual, because ionic charge transport is understood to be inhibited by Schottky-type charge transport barriers at the grain boundary (GB) regions, and both bulk and GB contributions are expected to appear in form of two separated semicircles in −Z′′ vs. Z′ plots.46 The reasons for the disappearance of the GB contributions are not entirely clear but the GB Schottky model may possibly not be valid here.
The ionic conductivity in the HTO sample is larger (the resistivity is lower) as compared to GTO, which is obvious from the different semicircle diameter (the difference is about a factor of 150). This large difference in resistivity is also evident in Fig. 12, where semi-logarithmic plots of resistivity ρ vs. reciprocal temperature 1/T are depicted. The resistivity values were obtained from the resistors in the equivalent circuits depicted in the insets of Fig. 11a and b. In the case of the HTO sample the open blue symbols represent resistivity values from impedance measurements under Ar atmosphere, which are almost identical to the ones from measurements under air (blue solid symbols). This implies that the ionic oxygen vacancy conduction in HTO may be independent from the ambient conditions.
The curve for the GTO ceramic pellet (red solid symbols) shows a deviation from the linear behaviour above ≈500 °C (773 K), which may again be associated with the unusual changes in the GTO lattice parameter at high T (Section 3.3).
The activation energy EA for the GTO ceramic was determined from the respective Arrhenius curves (lnρ vs. 1/T, not shown here) in the linear regime below ≈500 K (773 °C) only, whereas the entire T-range was considered for HTO. EA was found to be 0.79 eV and 0.88 eV for GTO and HTO respectively, which is in a typical range for oxygen ion conduction.49–54 Both EA values are rather similar, which suggests a similar charge transport mechanism. EA was determined by assuming a linear Arrhenius behaviour of ρ = ρ0
exp(EA/kBT). However, alternative models of ρ/T = ρ0
exp(EA/kBT) have also been proposed in the literature,55,56 in which case the activation energies would be slightly increased to 0.83 eV and 0.92 eV for GTO and HTO respectively. Within the second more advanced model the pre-exponential factors ρ0 have been claimed to exhibit a reciprocal relationship with the concentration of mobile species. From our fits we find ρ0 to be 1.3 10−1 Ω cm K−1 and 8.3 10−5 Ω cm K−1 for GTO and HTO respectively. The concomitant increased concentration of mobile species in HTO is in agreement with the lower nominal resistivity, which may be interpreted in terms of increased oxygen vacancy conduction in agreement with the increased oxygen vacancy disorder in HTO detected by Raman spectroscopy (Section 3.4). Since our ionic conductivity measurements were performed on GTO and HTO pellets sintered at 1400 °C, A and B anti-site defects may well be excluded in both compositions (see Table 1) and the increased oxygen vacancy conduction in HTO may imply a certain decoupling of anti-site and anion Frenkel defects.
Fig. 13 shows the capacitance values obtained from the equivalent circuit fits, plotted as dielectric permittivity ε′ vs. temperature. The ε′ values obtained from the bulk semicircles in GTO and HTO are both in a range that is typical for bulk values.46 The permittivity in the GTO is perceptibly increased as compared to the HTO, despite the equivalent crystal structure and similar cationic composition. The reasons for this increase are unclear, but we suspect that it may be correlated with the anomalous increase of the GTO unit cell (Section 3.3) and the significant changes in the Raman mode intensities (Section 3.4), both with increasing the heat treatment temperature. Although the HTO phase may be more relevant for applications due to an increased ionic conductivity, the GTO sample exhibits interesting unit cell volume changes accompanied by unusual Raman mode intensities, and both GTO and HTO species may deserve further research work in the future.
Both, the resistivity ρ and dielectric permittivity ε′ values presented in Fig. 12 and 13 were reproducible in control samples of GTO and HTO composition within an experimental error of ≈10%, which may arise from the uncertainty in measuring the pellet dimensions to calculate the specific parameters ρ and ε′.
This journal is © The Royal Society of Chemistry 2015 |