3. Results and Discussion
First, we discuss the crystal structure which will be a piece of information necessary for further discussion on the electrical and dielectric behavior of FeSbO
4. Herein, the powder X-ray diffraction pattern of FeSbO
4 was analyzed by conventional Rietveld refinement, as shown in
Figure 1. No trace of possible impurity from any secondary phase or unreacted precursor was observed. For the profile refinement, we assumed equal displacement factors for all atoms and a Thompson–Cox–Hastings profile of the reflections. The background was constructed by superposing Chebychev polynomials of a higher degree. A first attempt to refine the XRD patterns was made by imposing Fe
3+ and Sb
5+ randomly distributed in the 2a site and O occupying the 4f site and assuming the rutile-type structure (space group
P4
2/
mnm with a ≅ 4.63 Å and c ≅ 3.07 Å), which resulted in a reasonably high R
wp value (6.4%). However, the refinement readily converged to Bragg R-factors of ~2.6–2.9% (Rf-factor ~2.7–3.0%) as we assumed a trirutile-type (space group
P4
2/
mnm with a ≅ 4.63 Å and c ≅ 9.23 Å) with positional parameters such as Sb
5+ (at
z = y = z = 0 and
z = y = z = 1/2) and Fe
3+ (at
z = y = 0, z = 2/3, and
z = y = ½, z = 1/6), and equal probability distributed for Sb
5+ and Fe
3+ on (b) (at
z = y = 0, z = 1/3) and (at
z = y = 1/2, z = 5/6) sites. These metal sites are surrounded by O
2− octahedra, as reported for FeSbO
4 by Berry et al. [
10]. Then, we conclude by comparing both Rietveld refinements that the synthetic FeSbO
4 crystallizes in a superstructure of rutile, corresponding to a tripling along the axis of the rutile structure base (c’ = 3c) unit cell in which Fe
3+ ions are arranged in square planar layers separated by neighboring double layers of edge-connected SbO
6 octahedra. The refined structural parameters are summarized in
Table 1.
The X-ray density steadily decreases from ~6.1 to ~5.6 g∙cm
−3 with increasing powder calcination temperatures from 900 to 1000 °C, which is in good agreement with the density value of 5.82 g∙cm
−3 reported for the FeSbO
4 [
24]. The apparent densities and porosities of the ceramic bodies have also been calculated using the Archimedes immersion method. The highest relative density of (98.9 ± 0.2)% and lowest porosity of (1.1 ± 0.2)% were obtained for the C-900 ceramics pellets. However, the C-1000 ceramic exhibited the lowest relative density (68.2 ± 0.1)% and highest porosity (36.8 ± 0.1)%. Higher relative density and low porosity for C-900 ceramic were likely due to the smaller grain size of the calcined powder used for producing this ceramic. This indicates that the mean grain size range for the C-900 powder has a great influence on the kinetics of the sintering process.
The microstructures of the FeSbO
4 powder and sintered ceramics samples are presented in
Figure 2. It is clearly observed from the SEM micrographs, shown in
Figure 2a,c, that the grains have large lumps with irregular shapes and considerable agglomeration. It is found through high magnification that the agglomerates consist of primary particles with the size of (0.73 ± 0.51) μm (P-900) and (1.67 ± 0.54) μm (P-1000).
Figure 2b,d provide a summary of the SEM images of the fractured cross sections for the FeSbO
4 ceramic pallets sintered C-900 and C-1000, respectively. The grains have elongated and spherical structures, most likely associated with an open crystal structure of FeSbO
4. This structure contains two-dimensional layers of three-edge shared SbO
6 octahedra, which appear like an open honeycomb with holes, whereas FeO
6 octahedra connect each layer near holes, favoring low surface energy. We have also analyzed the micrographs of
Figure 2b,d using the intercept method, according to the ASTM E112 standard. The particle size distribution of the powder is given in
Figure 3a,b, which shows that C-900 ceramic pellet grain appears larger (~3.8 μm) and connected, resulting in a nearly fully dense ceramic with lower residual pores, which correlates with its high relative density (~99%) obtained by the Archimedes method. In contrast, the C-1000 ceramic shows smaller grains (~2.5 μm) and a low degree of sintering with a significant degree of porosity following the previously discussed result of ~39%. The higher interconnection grains number with clusters of small grains forming densely packed aggregates (2.55 ± 1.25) μm can also justify this significant porosity, as observed in the SEM image (
Figure 2d). Thus, we can see that the initial powder particle size distribution significantly affects the density and microstructure of the ceramics, as reported in other oxides [
25,
26,
27]. Herein, the larger particle sizes for the P-1000 powder resulted in ceramics with higher porosities than those produced using the P-900 powder with smaller particles. In fact, the higher surface energy for smaller particle sizes promotes a higher sintering rate of the compacted powder, resulting in an early packing process by favoring the dense aggregates to grow quickly, forming more prominent grains, and then inducing a significant decrease in the pore volume.
The EDS analysis reveals the existence of constituent elements, Fe, Sb, and O, confirming the purity of the FeSbO
4 powder and ceramics. For the C-900 sample, the actual atomic percentage was 21.1 ± 0.2 (Fe), 19.0 ± 0.2 (Sb), and 59.9 ± 1.3 (O) at.%, while 19.8 ± 0.3 (Fe), 17.3 ± 0.1 (Sb), and 62.9 ± 0.7 (O) at.% were observed for the C-1000 sample. Furthermore, these results also agree with the chemical composition of FeSbO
4 ceramics grains, whereas Fe:Sb atomic % ratios of (18.4 ± 0.2):(18.9 ± 0.1) and (19.6 ± 0.2):(17.6 ± 0.1) in the C-900 and C-1000 samples, respectively, are in the range of ideal tripuhyite (Fe:Sb = 1:1) [
15,
16].
Figure 4 shows the real part (
) of the complex electrical conductivity (
) versus frequency with varying temperatures from
to
for ceramic pellets uniaxially pressed at
and sintered at 1200 °C for 4 h. Analyzing the alternating current (AC) conductivity results in
Figure 4, both samples of FeSbO
4 (C-900 and C-1000) showed dispersion of conductivity,
, at high frequencies and dependence on temperature. In the direct current (DC) region with low frequencies (grain boundary region—GB) and
, the sample C-1000 showed a higher value in its conductivity in relation to the sample C-900. Both samples reveal a dependence with
according to Jeep Dyre’s model [
28], where
varies between 0.6 to 1.0 at high frequencies.
The C-900 sample with
showed a smaller dispersion in its imaginary conductivity (
) at low frequencies compared to sample C-1000
, according to the results in
Figure 5. At high temperatures, both samples showed excessive noise in the region between 0.1 to 10 Hz. For this reason, these measurements in this frequency range were omitted in
Figure 5. This problem is associated with noise in the sample holder electrodes in measurements performed at high temperatures.
In
Figure 6a,b the real (
) and imaginary (
) conductivities are presented as a function of frequency,
, performed at 25 °C. Analyzing the AC region for 𝑓 = 100 kHz, considering a grain contribution, sample C-1000 showed a lower conductivity than C-900; see
Table 2. Compared with the average grain size results obtained via SEM measurements (see
Figure 2), it can probably be associated with grain size. In other words, the small grains produce lower electrical conductivity,
. Millet et al. [
3] obtained a DC conductivity value of
for a FeSbO
4 green body pellet pressed at
Pa at room temperature (approximate value obtained graphically). However, at
, a substantial increase, around 1.9 times, in the ac electrical conductivity of the ceramic sample is observed, which has a
(C-1000) compared to the sample with
(C-900); see
Table 2.
According to Peko et al. [
29], certain concentrations of porosities possibly lodged in the grain boundary region favor an increase in this conductive region associated with a significant drop in the corresponding activation energy.
Figure 7 shows the frequency dependence of the room temperature dielectric constant of FeSbO
4 ceramics C-900 and C-1000 with 𝑓 = 100 kHz. The real dielectric constant was obtained from the capacitance data as a function of frequency, performed by the FRA Solatron 1260/Dielectric Bridge 1296. To convert the data from C to real dielectric constant, Equation (4) was used:
where
is the actual electrical permittivity (𝐹∙𝑚
−1),
is the capacitance (
),
𝑑 is the measurement cell thickness in 𝑚, 𝐴 is the area of the cylindrical face of the ceramic sample in
m2, and
.
Table 2 shows the results for the dielectric constant of the FeSbO
4 ceramic samples sintered at 1200 °C for 4 h at 𝑇 = 25 °C and 𝑓 = 100 kHz. According to
Table 2, the FeSbO
4 ceramic sample C-900 presented a dielectric constant 219.6% higher than the sample C-1000. All ceramics showed temperature-dependent dispersion in the dielectric constant and that this increase occurs at higher temperatures (
Figure 7). The fact that the dielectric constant was measured at high frequencies at
f = 100 kHz means that this increase in this magnitude,
, is possibly due to the increase in the average grain size and not due to the increase in porosity,
.
The complex impedance measurements,
and
, as a function of frequency with varying temperature were performed by applying an electrical potential of
and in a frequency range of
to
. The measurement method used was the two-probe technique. The resistivity results (
and
) in the complex plane were obtained from the normalization of the impedance data (
and
) through the geometric parameters of the sample:
given by the following Equations (5) and (6):
Several analysis models based on equivalent circuits are found in the literature, which is applied in different types of systems, especially in ceramics. In ceramic systems, where the relaxation frequency of the electrical impedance of the grain boundary is distinguishable from that of the grain, the most used equivalent circuit consists of two parallel resistor-capacitor blocks (RC) connected in series with each other. Among these analysis models that use equivalent circuits, the brick-layer model, so called because it treats the microstructure as an array of cubic grains separated by flat grain boundaries, is one of the most used for analyzing the electrical and dielectric properties of ceramic materials. The brick-layer model presupposes fine, continuous, and highly resistive grain boundaries, which would result in a one-dimensional current flow through the grain and transverse to the contour [
30,
31]. The complex electrical impedance associated with the parallel RC circuit of each region is described by Equation (7):
Multiplying the conjugate term
on both terms of Equation (7) and then separating the contribution from the real and imaginary parts of the complex impedance leads to Equation (8):
where the constant
is the relaxation time of the circuit and the real (
and imaginary
components of the complex impedance can be defined by Equations (9) and (10):
and
However, experimentally, impedance spectroscopy diagrams do not always present semicircles centered on the real axis. This decentralization results from the existence of a distribution of relaxation times,
and
, rather than a single value. In this case, an empirical correction is made in Equation (7) which takes the form shown in Equation (11) [
31]:
where the parameter, ψ, admits values between zero and one.
In this work, an electrical circuit composed of an electrode resistance (
) in series with two blocks that have resistances and capacitances of the grain regions (
and
) and grain boundary regions (
and
), where the complex impedance can be written by Equation (7):
Therefore, considering the decentralization factor in both terms of the complex impedance for the grain region and grain boundary of Equation (11), we then have the following relationship for the complex impedance of the iron antimonate ceramic given by Equation (13):
Figure 8a,b shows the Argand or Nyquist diagrams (the complex plane) for the electrical impedance spectrum of the samples that presented a single phase (C-900 and C-1000). Electrical measurements were performed at temperature levels ranging from 25 to 250 °C. In this diagram, each experimental point was measured at a frequency that increases from right to left between values of
to
. Analyzing the impedance curves for a temperature of 75 °C, sample C-900 presented semicircles with higher values on the abscissa than sample C-1000. In both ceramic samples, the semicircles decrease with the increasing temperature of the FeSbO
4 ceramic sample.
The resistances and capacitances of the grain regions (
and
) and grain boundary (
and
) were refined using the Zview software, version 2.9c, created by Dereck Johnson (North Carolina, USA) [
23] that models equivalent circuits using experimental data from impedance spectroscopy. Furthermore,
Figure 8 shows the Nyquist diagram of the samples studied in this work. The dashed red line represents the adjustment performed on the semicircles of the grain region and grain boundary of the FeSbO
4 ceramic samples. In addition, the deconvoluted grain and grain boundary regions (black dashed lines) are indicated.
From the adjustment of Equation (13) to the impedance spectroscopy experimental data, it was possible to determine the R and C values for the grain and grain boundary regions. To compare the electrical parameters, the resistances and capacitances were normalized by the geometric parameters (𝑑∙A
−1) of the sample and to obtain the activation energy of the grain and grain boundary regions through the Arrhenius plot [
27]; see
Figure 9. To obtain the grain activation, and grain boundary energies, the Arrhenius equation for electrical conductivity was used, given by Equation (14):
where
is the electrical conductivity associated with the charge carriers available for conduction in
,
is the activation energy required for the carrier to move in the crystal lattice of the ceramic in
,
is the Boltzmann constant in
, and
is the temperature in
. To obtain the activation energy,
, it was necessary to perform a transformation of ln
σ versus 1/𝑇, given by Equation (15):
Arrhenius plots of grain and grain boundary electrical conductivities for FeSbO
4 ceramics are shown in
Figure 9. Analyzing the Arrhenius graphs, it is possible to observe that the two ceramic samples of FeSbO
4 presented single activation energy, responsible for the AC electrical conduction mechanism both for the grain region and grain boundary region. This statement is valid only for the temperature range studied in this work, which was between 25 to 250 °C.
Analyzing
Figure 9b, it is possible to observe the influence of the calcination temperature (grain size) concerning the electrical conductivity of the grain, which is in orders of magnitude 2 from one sample to another. It is important to note that although the electrical conductivities of the grain boundary regions are smaller than the grain regions, their activation energies are very close, which is observed for both samples (
Table 3). According to Mariappan et al. [
32], even though the grain and grain boundary activation energy are virtually the same, they have a pre-exponential factor that differs considerably. Accordingly, in the grain boundary region, the blocking effect does not occur because of the high activation barriers but because of the geometric constriction effects. In addition to the activation energy involving AC electrical conduction processes, the activation energy at a frequency of 1 Hz was also determined, extrapolating as being a DC electrical conduction process; see
Figure 10. It is possible to observe single activation energy responsible for the DC electrical conduction process measured at low frequencies in both iron antimonate ceramic samples.
For knowledge, there are few experimental articles in the literature regarding the study of the electrical and dielectric properties of FeSbO
4 [
3,
4,
5,
6], particularly on the activation energy referring to the grain and grain boundary regions of the FeSbO
4 ceramic pellet. Millet et al. [
3] investigated the redox properties of pure V-doped FeSbO
4 through electrical conductivity measurements of a ceramic powder under mechanical pressure of 10
5 Pa, varying the temperature and with a controlled oxygen atmosphere. They determined the activation energy equal to 72.6 kJ∙mol
−1 and equivalent to 0.75 eV [
3]. Furthermore, Sing et al. [
4] through producing thin films of pure FeSbO
4, annealed at 450 °C for 2 h, and experimentally determined the activation energy around 0.49 eV from electrical resistance measurements as a function of temperature. Analyzing the results for the activation energy of the grain boundary region, as observed in
Table 3, in the case of a porous ceramic sample (C-1000), activation energy was measured,
, and an impact on its value was not observed to the detriment of the dense ceramic sample (C-900) with a value of
. Therefore, a direct influence of the porosity (P = 36.8 %) contained in this ceramic sample on the activation energy of the grain boundary region is not observed. Comparing the activation energies of the grain region (
and
) of the samples, C-900 and C-1000, respectively, it was observed that the ceramics, C-900 with larger average grain size,
, need a slightly higher activation energy to promote an increase in their AC electrical conductivity, corresponding to the contribution of the grain region. It can be concluded that despite the way of production of FeSbO
4 ceramics, pressed and sintered ceramics are different in relation to other works, only pressed ceramics or thin films; the results for activation energy obtained here in this manuscript are in good agreement with both works published by [
3,
4].
Figure 11 illustrates the loss factor (
) as a function of temperatures measured at 0.1, 1, and 100 kHz of the antimony ferrite samples. In the spectrum of
Figure 11a an expected increase in the loss factor as a function of temperature can be observed in all samples. The inset is a zoom for a temperature between 75 and 260 °C to distinguish the temperature-dependent loss factor measured at 100 kHz for the C-900 and C-1000 samples.
Figure 11b shows strong dispersion in loss factor at temperatures above 100 °C in the FeSbO
4 ceramic samples measured at 0.1 kHz to measurements performed at frequencies of 1 and 100 kHz. The understanding of this high dispersion at low frequencies can be explained mainly by the relationship between the electrical conductivity and the loss factor given by
where
is the frequency and,
is the real permittivity [
33]. As seen in
Table 2 and
Figure 11a, the porous ceramics (C-1000) have an increase in their conductivity due to the influence of pores in the grain boundary region; therefore, there was an increase in their loss factor,
. Emphasizing that this high dispersion at low frequencies may also have a smaller contribution as a function of grain size.
Considering that at high frequencies, there will only be the influence or contribution of the FeSbO
4 ceramic grain region, it is possible to observe in
Figure 11c and
Table 4 that the loss factor at the frequency of 100 kHz showed a slight increase as a function of the decrease of the average grain size. This hypothesis can be validated from an analysis of the location of the frequency equal to 10
5 Hz, as an example, in the complex impedance measurements shown in
Figure 8, performed at a temperature of 75 °C.
The sample C-1000 with a smaller grain size,
, showed a higher loss factor in the entire temperature range compared to the sample, C-900, with
. Analyzing the data contained in
Figure 11b, at a temperature of 25 °C, it was observed that the C-900 ceramic presented a loss factor equal to 1.20 and the C-1000 sample a factor equal to 2.03, at
= 1 kHz. However, at
, an increase in the loss factor was observed where samples C-900 and C-1000, respectively, presented values equal to 3.8 and 5.99, see
Table 4. It is concluded that the loss factor results corroborate the measurements of electrical conductivity, dielectric constant, and activation energy obtained in this work.