Next Article in Journal
Investigation on the Aerodynamic Performance of High-Pressure Cylinders under the Mode of Regulating Steam Supply with an Intermedium-Pressure Adjustment Valve
Previous Article in Journal
Development of a Novel Fabrication Process for Application in Glass Gas Electron Multiplier Detectors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrical Properties and Chemical Resistance of the Composites (1-x)Gd2Zr2O7·xMgO in Li-Containing Chloride Melts

1
Institute of High Temperature Electrochemistry of the Ural Branch of the Russian Academy of Sciences, 620990 Yekaterinburg, Russia
2
Institute of Metal Physics of the Ural Branch of the Russian Academy of Sciences, 620108 Yekaterinburg, Russia
3
Institute of Solid State Chemistry of the Ural Branch of the Russian Academy of Sciences, 620990 Yekaterinburg, Russia
*
Author to whom correspondence should be addressed.
Processes 2023, 11(4), 1217; https://doi.org/10.3390/pr11041217
Submission received: 28 February 2023 / Revised: 23 March 2023 / Accepted: 3 April 2023 / Published: 14 April 2023
(This article belongs to the Section Chemical Processes and Systems)

Abstract

:
Composites of (1-x)Gd2Zr2O7·xMgO were prepared by mixing gadolinium zirconate with freshly precipitated Mg(OH)2 followed by heat treatment at 1500 °C. Small concentrations of magnesium oxide dissolved in the complex oxide matrix of Gd2Zr2O7. This led to decrease in the lattice parameters of the matrix phase and a complex redistribution of Gd and Zr over the A and B sublattices. According to the impedance spectroscopy results of the studied samples, for (1-x)Gd2Zr2O7·xMgO (x = 0.05, 0.07, 0.10), the ionic conductivity was slightly higher than that for the undoped Gd2Zr2O7. The share of dominant ion transport did not change upon doping with magnesium oxide. The composites showed chemical resistance in a lithium halide (LiCl) melt and interacted with LiCl-xLi2O (x = 2 wt.%, 4 wt.%) melts at 650 °C with the formation of a Gd2O3 phase or a mixture of phases (Gd2O3, Li2ZrO3, ZrO2, LiGdO2, or LiGdCl2) on the ceramic surface, respectively.

1. Introduction

The pyrochemical processing of spent nuclear fuel is an effective way to reduce nuclear waste. Processes such as anodic dissolution, mass transfer, and the electrodeposition of actinides in molten salts have been studied in detail by scientists around the world [1,2,3,4]. The electrochemical reduction of SNF to a metallic form is the main stage of the electrorefining process for LiCl-xLi2O melts. As a result, actinide metal oxides are deposited on the cathode, while due to the proximity of electrode potentials, metallic lithium is formed on the cathode, which participates in the process of chemical reduction of actinide oxides. To obtain a high-quality product and avoid undesirable processes, such as the dissolution of the platinum anode at a low content of Li2O in the melt, i.e., ~0.25–0.50 wt.%, the lack of solubility of plutonium and americium oxides at high contents of Li2O, at ~3.4 wt.% and ~5.1 wt.%, respectively, or the degradation of structural materials, it is necessary to maintain a certain concentration of oxide ions in the working melt [5]. Therefore, devices for the continuous monitoring of the oxygen ion content in the working LiCl-xLi2O melt environment are needed.
Most of the ion-selective membrane sensors proposed for fabrication are made of ceramic materials, which must be inert in LiCl-based melts containing Li2O. Materials based on zirconium oxide used in oxide-halide melts are not stable enough to maintain stable operation for a long time. The formation of Li2ZrO3 is inevitable when using ZrO2 in LiCl-xLi2O melts, and with an increase in the content of Li2O, the stoichiometry of lithium zirconate changes to Li6Zr2O7 and Li8ZrO6 [6,7,8]. The coating of ceramics with a layer of lithium zirconate changes the conductivity from oxygen ion to cationic transport (by Li+-ions) [9,10,11,12,13,14]. The instability of electrolytes based on ZrO2 in the corrosive environment of Li+-containing melts [15,16,17] has stimulated the search for alternative materials—that is, sensors of the activity of O2− ions in melts.
One of the promising materials with high oxygen ionic conductivity is the Gd2Zr2O7 solid electrolyte with a pyrochlore structure. It is known that the structure of gadolinium zirconate withstands high doses of radiation due to the ability of the structure to endure cationic disorder. Under the action of radiation, the structure can undergo a structural phase transition from pyrochlore to a structure of defective fluorite, without being subjected to amorphization. In this regard, Gd2Zr2O7 is considered as a matrix for radioactive waste disposal [18,19,20,21,22]. Gadolinium zirconate is also used as a material for thermal barrier coatings [20,21,22], catalysts [23,24], solid electrolytes for SOFCs [25,26], and sensors [27].
The sensory activity of gadolinium zirconate has been insufficiently studied. However, there is evidence showing that when testing Ca+2-doped Gd2Zr2O7 as a NO2 sensor at 500 °C, the device showed the highest current value and the fastest response time, as well as the fastest recovery time. The results obtained were better than those of the commercial YSZ-based sensor [28].
In order to explore the possibility of using the material in the chloride melt as a sensor, it is necessary to study its chemical stability in the melt. For this purpose, it is necessary to determine the range of dopants that can ensure the acquisition of a high-density ceramic material with high, dominant oxygen ion conductivity. Previously, it was found that the use of lithium as a dopant for the Gd2Zr2O7 phase makes it possible to increase the oxygen ion conductivity but does not make it possible to obtain high-density ceramics [29,30]. Therefore, the problem of searching for dopants for the Gd2Zr2O7 phase remains relevant.
It has been established [31] that the homogeneous magnesium doping of Gd2Zr2O7 (solid-solution Gd2-xMgxZr2O7-x/2) can prevent the ion exchange of high-density ceramics with LiCl-xLi2O melts up to a Li2O concentration of 4 wt.%. Thus, a small concentration of magnesium is assumed to be able to protect the material from interaction with the melt [32]. For this article, materials based on Gd2Zr2O7 were chosen as the objects of study. In order to obtain high-density Gd2Zr2O7 ceramics, a composite with MgO additive was used. Composites based on gadolinium zirconate (1-x)Gd2Zr2O7·xMgO were investigated for their chemical resistance in oxide-chloride lithium melts for the first time. The structure and transport properties of (1-x)Gd2Zr2O7·xMgO were also studied for the first time.

2. Materials and Methods

Composites of (1-x)Gd2Zr2O7·xMgO were prepared by mixing gadolinium zirconate Gd2Zr2O7 with magnesium hydroxide. Gadolinium zirconate was synthesized by mixing with the starting reagents Gd2O3 (99.998%, VEKTON, RF) and ZrO2 (99.99%, REACHIM, RF). The oxide powders were ground together in a planetary ball mill (Fritsch Planetar Micro Mill Pulverisette 7, Fritsch GmbH, Federal Republic of Germany) with a rotation speed of 350 rpm in ethanol-exploiting zirconium balls for 8 h. After milling, the mixture was dried and annealed at 1200–1400 °C with steps of 100 °C. Freshly precipitated Mg(OH)2 was obtained from MgCO3 (99.99%, VEKTON, RF). The MgCO3 was dissolved in dilute nitric acid a with pH ~2–3, and then the resulting Mg(NO3)2 was caused to interact with a solution of NH4OH with the formation of a corresponding precipitate. The obtained Mg(OH)2 was dried and mixed with Gd2Zr2O7 in the amount required for the preparation of the samples of (1-x)Gd2Zr2O7·xMgO (x = 0.05, 0.07, 0.10). The corresponding calculations were carried out taking into account TG studies of the obtained Mg(OH)2. The resulting mixture was ground in the planetary ball mill at a rotation speed of 700 rpm for 2 h in ethanol. Then, the dried mixtures were pressed into cylindrical briquettes and treated at a temperature of 1500 °C for 12 h. The sintering of the cylindrical briquettes of undoped Gd2Zr2O7 was carried out at a temperature of 1550 °C for 15 h.
A crystal structure analysis of the samples was conducted using an X-ray diffractometer with Cu-Kα radiation (scanning rate of 0.5 °/min, steps of 0.01°) (Rigaku MiniFlex600, Japan) and the SmartLab Studio II (Rigaku, Japan) software with the PDF-2 2019 database (ICDD, USA). Silicon powder (NIST640f) was used as the external standard. The FullProf program was used for the refinement of the parameters with the Rietveld method.
Scanning electron microscopy of the sintered samples was conducted with Phenom ProX (Phenom-World, Holland). Before analysis, the surfaces of the samples were polishing on a diamond wheel. The composition of the samples was determined using energy-dispersive X-ray spectroscopy with a silicon drift detector and Phenom Pro Sute software.
The average density of the ceramics was determined using the geometric method. The value of the crystallographic density of the samples was taken as the standard.
The measurements of ac electrical conductivity, as a function of temperature and oxygen partial pressure, were made using the impedance spectroscopy method (Elins Z-1000P spectrometer, Elins Ltd., Moscow, Russia) in air over a temperature range of 300–1000 °C with cooling of 1 °/min and a pO2 range of 1 × 10–18–0.21 atm. The surfaces of the cylindrical-shaped samples were covered with platinum paste. The electrodes were burned in air at a temperature of 900 °C for 3 h. The ZView-4.0 software (Scribner Associate Inc., Southern Pines, NC, USA) was used to treat the impedance data. The oxygen partial pressure measurements were performed using an 8 mol.% Y2O3-stabilized zirconia (YSZ) sensor and an oxygen pump.
The stability of the ceramic composite samples was assessed in a glove box (SPEKS GB, Russia) in an argon atmosphere with an impurity oxygen content less than 2 ppm and a water content less than 1 ppm. The salts LiCl, LiCl-2 wt.% Li2O, and LiCl-4 wt.% Li2O were placed in a glassy carbon crucible. The salts LiCl (CAS No: 7447-41-8, Leverton Clarke Ltd., Basingstoke, UK) and Li2O (99.5 wt.%, Alfa Aesar, Haverhill, MA, USA) were used for the experiment. Before the experiment, the LiCl salt was subjected to additional purification by zone recrystallization. The salt in the crucible was heated in a shaft furnace at a temperature of 650 °C. The temperature in the working area was controlled using a Pt-Pt(Rh) thermocouple. The ceramics were preliminarily weighed on a balance and dried in an oven at 200 °C. They were then immersed in a melt of the corresponding salt. Exposure in the salt melts was carried out for 48 h. After holding in the melt, the samples were removed and washed in distilled water and ethanol. After drying the samples in an oven at 200 °C, the composites were analyzed by XRD and SEM.
The components of the melt were analyzed before the experiment and after 8, 16, 24, and 48 h by chemical analysis with inductively coupled plasma atomic emission spectroscopy (Optima 4300 DV Perkin Elmer spectrometer, USA).

3. Results

3.1. X-ray Analysis and SEM Characterization

The interpretation of the X-ray spectra (Figure 1) showed that the main phase in all the obtained composites was Gd2Zr2O7 with a pyrochlore structure (space group Fd-3m). The 0.90Gd2Zr2O7–0.10MgO composite contained a small amount of the second phase, which corresponded to the magnesium oxide. According to the XRD data, the samples with additives of MgO at the amounts of 5 wt.% and 7 wt.% were single-phase.
The lattice parameter decreased with the addition of MgO up to 7 wt.% (Figure 2a). However, the parameter remained unchanged over the concentration range of 7–10 wt.%.
The change in the lattice parameter indicates the formation of solid solution due to the dissolution of magnesium in the structure and the redistribution of cations between the A and B sublattices. The redistribution of cations with different ionic radii will affect the lattice parameter [33]. Figure 2b shows the X-ray pattern of the 0.95Gd2Zr2O7–0.05MgO sample processed using the Rietveld method in the FullProf program. In Figure 2b, one can see the diffraction data, which were treated using a gadolinium and zirconium redistributed model (Table 1). However, the inversion degree between the A and B sublattices grew from 8.15% (0.95Gd2Zr2O7–0.05MgO) to 12.7% (0.90Gd2Zr2O7–0.10MgO) with the increase in the Mg content.
The size of the Mg2+ ion is smaller than the Gd3+ ion in eightfold coordination [33]; thus, the parameter may decrease. In this regard, we assume that part of the Mg2+ dopes the Gd2Zr2O7 matrix and occupies the Gd3+ position (Table 2). Noting the misunderstanding of the exact amount of magnesium oxide included in this calculation, we considered the limited case, where the minimum concentration of the composite additive was included in the Gd2Zr2O7 structure. Accordingly, we can write the composition 0.95Gd2Zr2O7–0.05MgO as a solid solution: Gd1.88Mg0.05Zr0.07Zr1.93Gd0.07O7. The introduction of magnesium is accompanied by an increase in the degree of order of 0.95Gd2Zr2O7–0.05MgO by 11.4% compared to the undoped sample. In our previous work [31], it was shown that the introduction of magnesium into the structure could be achieved at an amount of 0.05 wt.%, and the calculations were conducted using the Rietveld method.
The calculations performed for the two different models of 0.95Gd2Zr2O7–0.05MgO sample presented in Table 1 and Table 2 show close results regarding the obtained calculation errors. It is probable that the lattice parameter is affected by a complex process associated with the disordering of the cationic and anionic sublattices, as well as the presence of small amounts of magnesium. Thus, to describe the structure of the obtained composites, one must not be limited to the X-ray method alone.
The SEM images confirm that all the (1-x)Gd2Zr2O7·xMgO samples contain impurities of MgO. Figure 3a–c shows densely sintered grains in the main phase, which tend to coalesce with each other with an increase in the concentration of MgO additive. Additionally, in all the SEM figures, there are dark spots of poorly crystallized (amorphous) phase. The EDS spectra show that such poorly crystallized spots belong to phase with an increased magnesium content (Figure 4a,b), and we believe that this is magnesium oxide. It is probable that due to amorphization, the low concentration of the composite additive MgO (x = 0.05 and x = 0.07) does not enable its identification in the composites by X-ray diffraction.
The relative densities of the obtained samples are all close to each other (Table 3), and there is a slight tendency for the relative density to increase with the content of magnesium oxide. It should be emphasized that higher temperatures and holding times are required to obtain high-density ceramics of undoped Gd2Zr2O7. Therefore, it can be concluded that the MgO additive contributes to an increase in the sinterability of the material, as shown in previous works [34,35].

3.2. Electrical Conductivities

Well-resolved impedance hodographs, obtained in the range from 300 °C to 1000 °C in air, are presented as two semicircles (Figure 5).
It was possible to separate the volume and grain boundary contributions to the total electrical conductivity. One semicircle, with a capacity of C = 10−11 F·cm−1, exceeds zero and is responsible for the bulk properties of the sample, and the second one, with a capacity of C = 10−9 F·cm−1, is responsible for the properties of the grain boundaries. By extrapolating the first semicircle to the OX axis, the bulk electrical conductivity was determined. The value of the bulk electrical conductivity of (1-x)Gd2Zr2O7-xMgO, shown in Figure 6, is slightly increased compared to that of the undoped Gd2Zr2O7 with a pyrochlore structure over the entire temperature range. The electrical conductivities of all the doped samples are close to each other. This effect is probably due to acceptor doping and the formation of oxygen vacancies.
The results obtained here were compared with the literature data on the conductivity of the most highly conductive phases [36,37]. Figure 7 shows the bulk conductivities of the materials with a pyrochlore structure. The electrical conductivity of the highly conductive phase Sm2-xCaxZr2O7-x/2 (x = 0, 0.05, 0.10) [36] is comparable with the data that we obtained for the composite (1-x)Gd2Zr2O7·xMgO. It can also be concluded (Figure 7a) that these samples have similar activation energies. Calcium doping of the gadolinium sublattice in Gd2Zr2O7 (Gd2-xCaxZr2O7-x/2 (x = 0.10)) significantly reduced the conductivity of the material. It can be seen in [37] that all the obtained samples have different activation energies of Gd2-xCaxZr2O7-x/2, and the electrical conductivity of the obtained compositions sharply decreases in the region of medium temperatures (Figure 7b). Thus, the addition of magnesium oxide MgO to gadolinium zirconate Gd2Zr2O7 did not lead to a decrease in conductivity; the values of ionic conductivity are comparable to those of the most highly conductive pyrochlores. At the same time, the addition of magnesium oxide made it possible to sinter the material at a lower temperature and obtain high-density ceramics. We believe that this method of obtaining high-density ceramics can be extended to other promising systems with a pyrochlore structure. Thus, for pyrochlores, magnesium oxide can be used both as an acceptor dopant and as a sintering additive.
The transport properties of the ceramics were studied in atmospheres with different oxygen partial pressures. The differentiation of the total conductivities into their ionic (σion) and electronic (σel) contributions was carried out through the analysis of the isotherms logσlogpO2. The total electrical conductivity (σtot) obtained in air at the oxygen partial pressure (pO2) = 0.21 atm is the sum of the ionic and electronic conductivities (1). The ionic electrical conductivity σion does not depend on pO2 (it is a plateau on the logσ–logpO2 graph).
σtot = σel + σion
Small additions of a dopant do not change the general nature of the dependence. The isotherms of conductivity are shown in Figure 8a–c, where the plateau of ionic conductivity can be observed over the entire temperature range. With an increase in the concentration of the heterogeneous additive to x = 0.10, the plots of electrical conductivity against pO2 show a positive slope in the region of high oxygen partial pressures, which indicates the appearance of a small p-type contribution to the total conductivity of the sample.
Based on the data obtained, the temperature dependence of ionic conductivity can be plotted. As can be seen in Figure 9, the ionic conductivity of all the composites is higher than that for the undoped phase. It can be assumed that this is due to an increase in oxygen vacancies as a result of acceptor doping and a complex redistribution of cations over the A and B sublattices.
The ionic transport number can be calculated as:
t i o n = σ i o n σ t o t
where σion is the ionic conductivity from the plateau of the logσpO2 dependencies between 10−18 atm and 10−4 atm, and σtot is the total conductivity at pO2 = 0.21 atm. As we can see from the equation, the ionic conductivity and the total conductivity increase, respectively, while the share (tion) does not change.
The ion transport numbers of the (1-x)Gd2Zr2O7·xMgO and undoped Gd2Zr2O7 samples are close (Figure 10). The transport numbers tend to increase with decreasing temperature and are close to unity over the entire temperature range. Thus, the addition of magnesium oxide did not affect the dominant ionic nature of transport of the Gd2Zr2O7 phase.

3.3. Chemical Resistance of LiCl and LiCl-Li2O Melts

Table 4 shows photographs of the (1-x)Gd2Zr2O7·xMgO ceramic samples before and after interaction with LiCl-xLi2O melts, where x = 0 wt%, 2 wt%, and 4 wt% at a temperature of 650 °C. As can be seen in the photo, with an increase in the concentration of lithium oxide in the melt, the degree of destruction of the samples increases. In particular, the 0.90Gd2Zr2O7-0.10MgO sample in a LiCl melt did not undergo destruction. The same sample in the LiCl-2 wt.% Li2O melt was partially destroyed and split into three parts, and in the LiCl-4 wt.% Li2O melt, it was completely destroyed and divided into three large parts and many small parts, which also underwent dissolution, judging by the presence of a white coating on the surface of the electrolyte during the experiment and chemical analysis. High-density ceramic samples of Gd2Zr2O7 were examined in our earlier work [27]; however, it was shown that Gd2Zr2O7 interacted with the LiCl-2 wt.% Li2O melt during high-temperature exposure.
Table 5 shows the change in mass of the ceramics after the exposure of the ceramic samples of (1-x)Gd2Zr2O7·xMgO in LiCl-xLi2O melts at a temperature of 650 °C. After holding the samples in the LiCl melt, there was no significant change in the density of the ceramics and no mass loss. Despite the fact that the 0.93Gd2Zr2O7-0.07MgO ceramics cracked, the mass loss during the study was 0.01%. It is likely that the presence of small defects in the form of cracks in 0.93Gd2Zr2O7-0.07MgO significantly affected the strength characteristics of the material and led to destruction; thus, this fact should be taken into account in further studies. When 2 wt.% Li2O was added to the melt, the 0.90Gd2Zr2O7-0.10MgO ceramics were destroyed, as a result of which the mass loss was ~3%. In the melt with 4 wt.% Li2O, there was a significant mass loss in all the ceramics under study, decreasing from ~3% to ~11.5%. The greatest loss occurred in the 0.90Gd2Zr2O7-0.10MgO sample.
The X-ray phase analysis data presented in Figure 11a show that the 0.95Gd2Zr2O7-0.05MgO sample interacted with the LiCl melt, since during the experiment, a phase of gadolinium oxide formed on the surface of the material. The samples (0.93Gd2Zr2O7-0.07MgO, 0.90Gd2Zr2O7-0.10MgO) did not chemically interact with the working medium during the 48 h of exposure. Figure 11b shows that in a more aggressive melt containing lithium oxide at an amount of 2 wt.%, exposure led to the interaction of all the ceramics with the melt, together with the formation of a gadolinium oxide phase.
The concentration of lithium oxide also significantly affects the rate of destruction of the material and the formation of end products through the chemical interaction of the material with the melt. Figure 12 shows the X-ray diffraction patterns of the composite surface after 48 h of interaction with the LiCl-4 wt.% Li2O melt. The surface consists of a mixture of phases formed by the chemical interaction of gadolinium zirconate with the melt components. Such a process can be described by Equations (3)–(6):
Gd 2 Zr 2 O 7 + 2 Li 2 O = 2 Li 2 ZrO 3 + Gd 2 O 3
Gd 2 O 3 + Li 2 O = 2 LiGdO 2
LiGdO 2 + 4 LiCl = 2 LiGdCl 4 + 2 Li 2 O
or
Gd 2 O 3 + 8 LiCl = 2 LiGdCl 4 + 3 Li 2 O
The SEM data of the composites are confirmed by the XRD data. Figure 13a–c shows the surfaces of composites exposed in a LiCl melt at a temperature of 650 °C. It was found that the surface of 0.95Gd2Zr2O7-0.05MgO had grains of different contrasts. The surfaces of the remaining compositions remained identical to the initial state. The surfaces of the composites exposed in LiCl-2 wt.% Li2O were distinguished by grains with different morphologies and contrasts which were unevenly distributed over the ceramic surfaces (Figure 13d–f). The smallest grain growth and grain concentration were recorded for the composite 0.93Gd2Zr2O7-0.07MgO. EDS spectra confirmed (Figure 14) that the dark surface grains corresponded to a phase with a high content of zirconium, and the light grains corresponded to a phase with a high content of gadolinium. With an increase in the content of lithium oxide to 4 wt.%, the morphology of the grains also changed (Figure 13g–i), and their enlargement occurred.
The chemical analysis of the (1-x)Gd2Zr2O7·xMgO composite did not establish the presence of Gd or Zr ceramic components in the LiCl melt. With an increase in the concentration of lithium oxide in the melt, the concentrations of zirconium and gadolinium as components of the melt increased (Figure 15). Additionally, depending on the composition of the ceramics, the saturation time of the melt, with respect to gadolinium, varied. At x = 0.05, the saturation time for the LiCl-4 wt.% Li2O melt was 8 h, and at x = 0.07 and x = 0.10, the saturation times were 24 and 30 h, respectively.
Thus, the results obtained showed that the addition of magnesium oxide did not lead to an increase in the chemical stability of the phase Gd2Zr2O7 in the melts of LiCl-Li2O. We believe that this is due to the morphology of the obtained composites when the magnesium oxide phase was localized in the form of individual grains. It is probably that the situation would be more favorable with the addition of magnesium oxide, encapsulating the grains of the main phase, Gd2Zr2O7, and preventing interaction with the melt. Since the system under study was demonstrated to be capable of obtaining high oxygen ion conductivities and high-density ceramics, it is of interest for further research, for example, using higher concentrations of magnesium oxide or changing the method for obtaining a composite. This will be the topic of our further research.
It should be emphasized that there are few studies on the chemical resistance of ceramics in halide melts. The investigations are mainly carried out for simple oxides (doped with an acceptor impurity), for example: ZrO2(Y2O3), ZrO2(MgO), ZrO2(Sc2O3), ZrO2(CaO), CeO2(Gd2O3) [38,39]. It was shown that the increase in the content of Li2O led to the destruction of the samples. Therefore, investigations of Gd2Zr2O7-based pyrochlores are the most promising.

4. Conclusions

Composites with the general formula (1-x)Gd2Zr2O7·xMgO were obtained. Small concentrations of magnesium oxide were dissolved in the complex oxide matrix of Gd2Zr2O7 with the formation of a solid solution, which manifested itself through a decrease in the lattice parameters of the matrix phase. This led to the formation of oxygen vacancies Processes 11 01217 i013 and a complex redistribution of the Gd and Zr over the A and B sublattices. The ionic conductivity of the studied samples of (1-x)Gd2Zr2O7·xMgO (x = 0.05, 0.07, 0.10) was slightly higher than that of the undoped phase. The oxygen ionic conductivity reached 6.6 × 10−3 Ohm·cm at 800 °C and 8.1 × 10−4 Ohm·cm at 600 °C for (1-x)Gd2Zr2O7·xMgO (x = 0.07). The transfer numbers, below 500 °C, reached 100%. The share of dominant ion transport did not change upon doping with magnesium oxide.
Composites based on gadolinium zirconate were investigated for their chemical resistance in oxide-chloride lithium melts for the first time. Composites of (1-x)Gd2Zr2O7·xMgO (x = 0.07, 0.10) showed chemical resistance in lithium halide (LiCl) melts for 48 h at 650 °C. However, when Li2O was introduced into the melt at the amounts of 2 wt.% and 4 wt.%, the composites chemically interacted with the melt component (Li2O), with the formation of a Gd2O3 phase or a mixture of phases (Gd2O3, Li2ZrO3, ZrO2, LiGdO2, or LiGdCl2) on the ceramic surface, respectively. This interaction was also confirmed by the chemical analysis of the melt after the experiments.
Additives of magnesium oxide had a significant effect on sintering, which ensured the production of high-density samples at lower temperatures and shorter heat treatment times.

Author Contributions

Conceptualization, I.A. (Irina Animitsa); methodology, Y.Z. and I.A. (Irina Anokhina); X-ray diffraction analysis, V.V. and N.K.; melt research and methodology, M.E.; writing—original draft preparation, I.A. (Irina Animitsa) and I.A. (Irina Anokhina); writing—review and editing, I.A. (Irina Animitsa) and I.A. (Irina Anokhina). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Choi, E.-Y.; Hur, J.-M.; Choi, I.-K.; Kwon, S.G.; Kang, D.-S.; Hong, S.S.; Shin, H.-S.; Yoo, M.A.; Jeong, S.M. Electrochemical reduction of porous 17 kg uranium oxide pellets by selection of an optimal cathode/anode surface area ratio. J. Nucl. Mater. 2011, 418, 87–92. [Google Scholar] [CrossRef]
  2. Willit, J.L.; Miller, W.E.; Battles, J.E. Electrorefining of uranium and plutonium—A literature review. J. Nucl. Mater. 1992, 195, 229–249. [Google Scholar] [CrossRef]
  3. Kurata, M.; Inoue, T.; Serp, J.; Ougier, M.; Glatz, J.-P. Electro-chemical reduction of MOX in LiCl. J. Nucl. Mater. 2004, 328, 97–102. [Google Scholar] [CrossRef]
  4. Merwin, A.; Williamson, M.A.; Willit, J.L.; Chidambaram, D. Review—Metallic Lithium and the Reduction of Actinide Oxides. J. Electrochem. Soc. 2017, 164, H5236–H5246. [Google Scholar] [CrossRef]
  5. Choi, E.-Y.; Choi, I.-K.; Hur, J.-M.; Kang, D.-S.; Shin, H.-S.; Jeong, S.M. In Situ Electrochemical Measurement of O2− Concentration in Molten Li2O/LiCl during Uranium Oxide Reduction. J. Solid State Chem. 2012, 15, E11–E13. [Google Scholar] [CrossRef]
  6. Sakamura, Y.; Iizuka, M.; Kitawaki, S.; Nakayoshi, A.; Kofuji, H. Formation and reduction behaviors of zirconium oxide compounds in LiCl-Li2O melt at 923 K. J. Nucl. Mater. 2015, 466, 269–279. [Google Scholar] [CrossRef]
  7. Cho, S.-H.; Kim, D.-Y.; Kwon, S.; Yoon, B.-H.; Lee, J.-H. High-temperature corrosion characteristics of yttria-stabilized zirconia material in molten salts of LiCl-Li2O and LiCl-Li2O-Li. J. Nucl. Sci. Technol. 2018, 55, 97–103. [Google Scholar] [CrossRef]
  8. Pfeiffer, H.; Knowles, K.M. Reaction mechanisms and kinetics of the synthesis and decomposition of lithium metazirconate through solid-state reaction. J. Eur. Ceram. Soc. 2004, 24, 2433–2443. [Google Scholar] [CrossRef]
  9. Pantyukhina, M.I.; Andreev, O.L.; Martem’yanova, Z.S.; Batalov, N.N. Cation Conductivity of Li8ZrO6–LiYO2 Solid Solutions. Inorg. Mater. 2004, 40, 404–406. [Google Scholar] [CrossRef]
  10. Sherstobitova, E.A.; Gubkin, A.F.; Bobrikov, I.A.; Kalashnova, A.V.; Pantyukhina, M.I. Bottle-necked ionic transport in Li2ZrO3: High temperature neutron diffraction and impedance spectroscopy. Electrochim. Acta 2016, 209, 574–581. [Google Scholar] [CrossRef]
  11. Hellstrom, E.E.; Gool, W.V. Li ion conduction in Li2ZrO3, Li4ZrO4, and LiScO2. Solid State Ion. 1981, 2, 59–64. [Google Scholar] [CrossRef]
  12. Zhan, X.; Cheng, Y.-T.; Shirpour, M. Nonstoichiometry and Li-Ion Transport in Lithium Zirconate: The Role of Oxygen Vacancies. J. Am. Ceram. Soc. 2018, 101, 4053–4065. [Google Scholar] [CrossRef]
  13. Pantyukhina, M.I.; Obrosov, V.P.; Stepanov, A.P.; Voronin, V.I.; Batalov, N.N. Study of Ion Transport in Li2ZrO3 Solid Electrolytes with Different Lithium Isotope Ratios. Crystallogr. Rep. 2004, 49, 676–679. [Google Scholar] [CrossRef]
  14. Wyers, G.P.; Cordfunke, E.H.P. Phase relations in the system Li2O-ZrO2. J. Nucl. Mater. 1989, 168, 24–30. [Google Scholar] [CrossRef]
  15. Shishkin, A.V.; Shishkin, V.Y.; Nikolaev, A.Y.; Pankratov, A.A.; Suzdaltsev, A.V.; Zaikov, Y.P. Reduction of ZrO2 in LiCl-Li2O Melt during Electrolysis. J. Electrochem. Soc. 2022, 169, 116506. [Google Scholar] [CrossRef]
  16. Choi, E.-Y.; Meon, M.K.; Kim, S.-W.; Paek, S. Dilution of Li-Li2O in a metallic fuel produced through oxide reduction using ZrO2-assisted rinsing in molten LiC. J. Nucl. Mater. 2020, 533, 152107. [Google Scholar] [CrossRef]
  17. Nikolaev, A.; Suzdaltsev, A.; Pavlenko, O.; Zaikov, Y.; Kurennykh, T.; Vykhodets, V. Reduction of ZrO2 during SNF Pyrochemical Reprocessing. J. Electrochem. Soc. 2021, 168, 036506. [Google Scholar] [CrossRef]
  18. Chen, Q.-Y.; Shih, K.-M.; Meng, C.-M.; Liao, C.-Z.; Wang, L.-L.; Xie, H.; Lv, H.-Y.; Wu, T.; Ji, S.-Y.; Huang, Y.-Z. Ab initio density functional theory study of uranium solubility in Gd2Zr2O7 pyrochlores. Chin. Phys. C 2015, 1–20. [Google Scholar] [CrossRef]
  19. Shu, X.; Qing, Q.; Li, B.; Yang, Y.; Li, L.; Zhang, H.; Shao, D.; Lu, X. Rapid immobilization of complex simulated radionuclides by as-prepared Gd2Zr2O7 ceramics without structural design. J. Nucl. Mater. 2019, 526, 151782. [Google Scholar] [CrossRef]
  20. Lang, M.; Zhang, F.; Zhang, J.; Wang, J.; Lian, J.; Weber, W.J.; Schuster, B.; Trautmann, C.; Neumann, R.; Ewing, R.C. Review of A2B2O7 pyrochlore response to irradiation and pressure. Nucl. Instrum. Methods Phys. Res. B Beam Interact. Mater. At. 2010, 268, 2951–2959. [Google Scholar] [CrossRef]
  21. Wang, J.; Wang, J.; Zhang, Y.; Li, Y.; Teng, Y.; Wang, Z.; Tan, H. Flux synthesis and chemical stability of Nd and Ce co-doped (Gd1-xNdx)2(Zr1-xCex)2O7 (0 ≤ x ≤ 1) pyrochlore ceramics for nuclear waste forms. Ceram. Int. 2017, 43, 17064–17070. [Google Scholar] [CrossRef]
  22. Chung, C.K.; Shamblin, J.; O’Quinn, E.C.; Shelyug, A.; Gussev, I.; Lang, M.; Navrotsky, A. Thermodynamic and structural evolution of Dy2Ti2O7 pyrochlore after swift heavy ion irradiation. Acta Mater. 2018, 145, 227–234. [Google Scholar] [CrossRef]
  23. Jitta, R.R.; Gundeboina, R.; Veldurthi, N.K.; Guje, R.; Muga, V. Defect pyrochlore oxides: As photocatalyst materials for environmental and energy applications—A review. J. Chem. Technol. Biotechnol. 2015, 90, 1937–1948. [Google Scholar] [CrossRef]
  24. Zinatloo-Ajabshir, S.; Salavati-Niasari, M.; Zinatloo-Ajabshir, Z. Nd2Zr2O7-Nd2O3 nanocomposites: New facile synthesis, characterization and investigation of photocatalytic behavior. Mater. Lett. 2016, 180, 27–30. [Google Scholar] [CrossRef]
  25. Radhakrishnan, A.N.; Rao, P.P.; Mahesh, S.K.; Vaisakhan Thamp, D.S.; Koshy, P. Role of bond strength on the lattice thermal expansion and oxide ion conductivity in quaternary pyrochlore solid solutions. Inorg. Chem. 2012, 51, 2409–2419. [Google Scholar] [CrossRef] [PubMed]
  26. Shlyakhtina, A.; Shcherbakova, L.G. New solid electrolytes of the pyrochlore family. Russ. J. Electrochem. 2012, 48, 1–25. [Google Scholar] [CrossRef]
  27. Pavlenko, O.B.; Anokhina, I.A.; Dedyukhin, A.E.; Voronin, V.I.; Nikolaev, A.Y.; Mazannikov, M.V.; Nikitina, E.V.; Karfidov, E.A.; Kholkina, A.S.; Zaikov, Y.P. Interaction between gadolinium zirconate and LiCl–Li2O melt. Ceram. Int. 2021, 47, 20151–20160. [Google Scholar] [CrossRef]
  28. Zhong, F.; Zhao, J.; Shi, L.; Xiao, Y.; Cai, G.; Zheng, Y.; Long, J. Alkaline-Earth Metals-Doped Pyrochlore Gd2Zr2O7 as Oxygen Conductors for Improved NO2 Sensing Performance. Sci. Rep. 2017, 7, 4684. [Google Scholar] [CrossRef] [Green Version]
  29. Anokhina, I.A.; Animitsa, I.E.; Buzina, A.F.; Nokhrin, S.S.; Zaikov, Y.P.; Voronin, V.I.; Vykhodets, V.B.; Kurennykh, T.E.; Kazakova, V.N. Electrical properties of Li+-substituted solid solutions based on Gd2Zr2O7. Russ. J. Phys. Chem. A 2021, 95, 2426–2431. [Google Scholar] [CrossRef]
  30. Anokhina, I.A.; Animitsa, I.E.; Voronin, V.I.; Vykhodets, V.B.; Kurennykh, T.E.; Molchanova, N.G.; Vylkov, A.I.; Dedyukhin, A.E.; Zaikov, Y.P. The structure and electrical properties of lithium doped pyrochlore Gd2Zr2O7. Ceram. Int. 2021, 47, 1949–1961. [Google Scholar] [CrossRef]
  31. Anokhina, I.; Pavlenko, O.; Proskurnina, N.; Dedyukhin, A.; Animitsa, I. The Gd2-xMgxZr2O7-x/2 solid solution: Ionic conductivity and chemical stability in the melt of LiCl-Li2O. Materials 2022, 15, 4079. [Google Scholar] [CrossRef] [PubMed]
  32. Golosov, O.A.; Khvostov, S.S.; Glushkova, N.V.; Evseev, M.V.; Staritsyn, S.V.; Zaikov, Y.P.; Kovrov, V.A.; Nikitina, E.V.; Kholkina, A.S.; Kazakovtseva, N.A.; et al. Corrosive and mechanical resistance of MgO ceramics under metallizing and mild chlorination of spent nuclear fuel in molten salts. Ceram. Int. 2021, 47, 3306–3311. [Google Scholar] [CrossRef]
  33. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  34. Qu, Z.; Wan, C.; Pan, W. Thermal expansion and defect chemistry of MgO-doped Sm2Zr2O7. Chem. Mater. 2007, 19, 4913–4918. [Google Scholar] [CrossRef]
  35. Liu, Z.-G.; Ouyang, J.-H.; Sun, K.-N.; Zhou, Y. Influence of magnesia doping on structure and electrical conductivity of pyro-chlore type GdSmZr2O7. Adv. Appl. Ceram. 2012, 111, 214–219. [Google Scholar] [CrossRef]
  36. Shlyakhtina, A.V.; Abrantes, J.C.C.; Gomes, E.; Lyskov, N.V.; Konysheva, E.Y.; Kharitonova, E.P.; Karyagina, O.K.; Kolbanev, I.V.; Shcherbakova, L.G. Evolution of oxygen-ion and proton conductivity in Ca doped Ln2Zr2O7 (Ln = Sm, Gd) zirconates, located near pyrochlore-fluorite phase boundary. Materials 2019, 12, 2452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Jiang, L.; Wang, C.; Wang, J.; Liu, F.; You, R.; Lv, S.; Zeng, G.; Yang, Z.; He, J.; Liu, A.; et al. Pyrochlore Ca-doped Gd2Zr2O7 solid state electrolyte type sensor coupled with ZnO sensing electrode for sensitive detection of HCHO. Sens. Actuators B Chem. 2020, 309, 127768. [Google Scholar] [CrossRef]
  38. Cho, S.H.; Park, S.B.; Lee, J.H.; Hur, J.M.; Lee, H.S. Hot corrosion behavior of ZrO2–MgO coatings in LiCl–Li2O molten salt. Mater. Chem. Phys. 2012, 131, 743–751. [Google Scholar] [CrossRef]
  39. Valtseva, A.I.; Pershin, P.S.; Suzdaltsev, A.V.; Zaikov, Y.P. Research of oxygen-conducting ceramic materials for lithium chloride melt in reactors for pyrochemical processing of spent nuclear fuel. J. Phys. Conf. Ser. 2020, 1683, 032029. [Google Scholar] [CrossRef]
Figure 1. X-ray patterns obtained in air after annealing of the composites at 1500 °C: (1-x) Gd2Zr2O7·xMgO. The black dots denote the MgO phase reflections.
Figure 1. X-ray patterns obtained in air after annealing of the composites at 1500 °C: (1-x) Gd2Zr2O7·xMgO. The black dots denote the MgO phase reflections.
Processes 11 01217 g001
Figure 2. (a) Graph of the dependence of the lattice parameter on the composition. (b) X-ray diffraction pattern of the 0.95Gd2Zr2O7–0.05MgO sample processed using the FullProf program.
Figure 2. (a) Graph of the dependence of the lattice parameter on the composition. (b) X-ray diffraction pattern of the 0.95Gd2Zr2O7–0.05MgO sample processed using the FullProf program.
Processes 11 01217 g002
Figure 3. SEM images of (1-x)Gd2Zr2O7·xMgO samples at (a) x = 0.05; (b) x = 0.07; (c) x = 0.10.
Figure 3. SEM images of (1-x)Gd2Zr2O7·xMgO samples at (a) x = 0.05; (b) x = 0.07; (c) x = 0.10.
Processes 11 01217 g003
Figure 4. (a) SEM images of 0.90Gd2Zr2O7–0.10MgO. (b) EDS spectra of 0.90Gd2Zr2O7–0.10MgO in point 3.
Figure 4. (a) SEM images of 0.90Gd2Zr2O7–0.10MgO. (b) EDS spectra of 0.90Gd2Zr2O7–0.10MgO in point 3.
Processes 11 01217 g004
Figure 5. Impedance hodographs of 0.95Gd2Zr2O7-0.05MgO obtained in air in the temperature ranges of (a) 300–373 °C, (b) 497–595 °C, and (c) 650–799 °C.
Figure 5. Impedance hodographs of 0.95Gd2Zr2O7-0.05MgO obtained in air in the temperature ranges of (a) 300–373 °C, (b) 497–595 °C, and (c) 650–799 °C.
Processes 11 01217 g005
Figure 6. Temperature dependence of bulk electrical conductivity for the (1-x)Gd2Zr2O7·xMgO composites.
Figure 6. Temperature dependence of bulk electrical conductivity for the (1-x)Gd2Zr2O7·xMgO composites.
Processes 11 01217 g006
Figure 7. Temperature dependence of bulk electrical conductivity for the (1-x)Gd2Zr2O7·xMgO composites, Sm2-xCaxZr2O7-x/2 (x = 0, 0.05, 0.10) [36], and Gd2-xCaxZr2O7-x/2 (x = 0.10) [36]. (a) Temperature dependence of bulk electrical conductivity for the (1-x)Gd2Zr2O7·xMgO composites and Gd2-xCaxZr2O7-x/2 (x = 0, 0.02, 0.05, 0.10) [37] (b).
Figure 7. Temperature dependence of bulk electrical conductivity for the (1-x)Gd2Zr2O7·xMgO composites, Sm2-xCaxZr2O7-x/2 (x = 0, 0.05, 0.10) [36], and Gd2-xCaxZr2O7-x/2 (x = 0.10) [36]. (a) Temperature dependence of bulk electrical conductivity for the (1-x)Gd2Zr2O7·xMgO composites and Gd2-xCaxZr2O7-x/2 (x = 0, 0.02, 0.05, 0.10) [37] (b).
Processes 11 01217 g007
Figure 8. Dependence of the electrical conductivity of the composition on the oxygen partial pressure for (a) 0.95Gd2Zr2O7-0.05MgO, (b) 0.93Gd2Zr2O7-0.07MgO, (c) 0.90Gd2Zr2O7-0.10MgO.
Figure 8. Dependence of the electrical conductivity of the composition on the oxygen partial pressure for (a) 0.95Gd2Zr2O7-0.05MgO, (b) 0.93Gd2Zr2O7-0.07MgO, (c) 0.90Gd2Zr2O7-0.10MgO.
Processes 11 01217 g008
Figure 9. Temperature dependence of ionic conductivity (a). Dependence of ionic conductivity on x in (1-x)Gd2Zr2O7·xMgO at 600 °C (b).
Figure 9. Temperature dependence of ionic conductivity (a). Dependence of ionic conductivity on x in (1-x)Gd2Zr2O7·xMgO at 600 °C (b).
Processes 11 01217 g009
Figure 10. The values of the ion transfer numbers for (1-x)Gd2Zr2O7·xMgO (x = 0, 0.05, 0.07, 0.10).
Figure 10. The values of the ion transfer numbers for (1-x)Gd2Zr2O7·xMgO (x = 0, 0.05, 0.07, 0.10).
Processes 11 01217 g010
Figure 11. X-ray patterns of (1-x)Gd2Zr2O7·xMgO after exposure at T = 650 °C in (a) LiCl melt and (b) LiCl-2 wt.% Li2O melt.
Figure 11. X-ray patterns of (1-x)Gd2Zr2O7·xMgO after exposure at T = 650 °C in (a) LiCl melt and (b) LiCl-2 wt.% Li2O melt.
Processes 11 01217 g011
Figure 12. X-ray patterns of (1-x)Gd2Zr2O7·xMgO ceramics after exposure in LiCl-4 wt% Li2O melt at T = 650 °C. Asterisks—Gd2Zr2O7, black diamonds—Li2ZrO3, black circles—Gd2O3, white diamonds—ZrO2, white circles—LiGdO2, crosses—LiGdCl4.
Figure 12. X-ray patterns of (1-x)Gd2Zr2O7·xMgO ceramics after exposure in LiCl-4 wt% Li2O melt at T = 650 °C. Asterisks—Gd2Zr2O7, black diamonds—Li2ZrO3, black circles—Gd2O3, white diamonds—ZrO2, white circles—LiGdO2, crosses—LiGdCl4.
Processes 11 01217 g012
Figure 13. Scanning electron microscopy images of (a) 0.95Gd2Zr2O7-0.05MgO, (b) 0.93Gd2Zr2O7-0.07MgO, and (c) 0.90Gd2Zr2O7-0.10MgO after exposure for 48 h in LiCl at a T = 650 °C. Scanning electron microscopy images of (d) 0.95Gd2Zr2O7-0.05MgO, (e) 0.93Gd2Zr2O7-0.07MgO, and (f) 0.90Gd2Zr2O7-0.10MgO after exposure for 48 h in a LiCl-2 wt.% Li2O melt at T = 650 °C. Scanning electron microscopy images of (g) 0.95Gd2Zr2O7-0.05MgO, (h) 0.93Gd2Zr2O7-0.07MgO, and (i) 0.90Gd2Zr2O7-0.10MgO after exposure for 48 h in a LiCl-4 wt.% Li2O melt at T = 650 °C.
Figure 13. Scanning electron microscopy images of (a) 0.95Gd2Zr2O7-0.05MgO, (b) 0.93Gd2Zr2O7-0.07MgO, and (c) 0.90Gd2Zr2O7-0.10MgO after exposure for 48 h in LiCl at a T = 650 °C. Scanning electron microscopy images of (d) 0.95Gd2Zr2O7-0.05MgO, (e) 0.93Gd2Zr2O7-0.07MgO, and (f) 0.90Gd2Zr2O7-0.10MgO after exposure for 48 h in a LiCl-2 wt.% Li2O melt at T = 650 °C. Scanning electron microscopy images of (g) 0.95Gd2Zr2O7-0.05MgO, (h) 0.93Gd2Zr2O7-0.07MgO, and (i) 0.90Gd2Zr2O7-0.10MgO after exposure for 48 h in a LiCl-4 wt.% Li2O melt at T = 650 °C.
Processes 11 01217 g013
Figure 14. SEM images and EDS data for the sample of 0.90Gd2Zr2O7-0.10MgO after exposure for 48 h in a LiCl-2 wt.% Li2O melt at T = 650 °C (a). SEM images of and EDS spectrum of the 0.90Gd2Zr2O7-0.10MgO sample after exposure for 48 h in the LiCl-2 wt.% Li2O melt at T = 650 °C (b).
Figure 14. SEM images and EDS data for the sample of 0.90Gd2Zr2O7-0.10MgO after exposure for 48 h in a LiCl-2 wt.% Li2O melt at T = 650 °C (a). SEM images of and EDS spectrum of the 0.90Gd2Zr2O7-0.10MgO sample after exposure for 48 h in the LiCl-2 wt.% Li2O melt at T = 650 °C (b).
Processes 11 01217 g014
Figure 15. (a) Chemical analysis data on the content of Gd in the melts after 48 h of exposure for (1-x)Gd2Zr2O7·xMgO. (b) Chemical analysis data on the Zr content in the melts after 48 h of holding for (1-x)Gd2Zr2O7·xMgO.
Figure 15. (a) Chemical analysis data on the content of Gd in the melts after 48 h of exposure for (1-x)Gd2Zr2O7·xMgO. (b) Chemical analysis data on the Zr content in the melts after 48 h of holding for (1-x)Gd2Zr2O7·xMgO.
Processes 11 01217 g015
Table 1. The refined coordinates of atoms of 0.95Gd2Zr2O7–0.05MgO in a cubic unit cell (space group Fd-3m) in the Gd and Zr disorder model.
Table 1. The refined coordinates of atoms of 0.95Gd2Zr2O7–0.05MgO in a cubic unit cell (space group Fd-3m) in the Gd and Zr disorder model.
AtomWyckhoff Symbolx/ay/bz/cOccupancy
Gd16c0.0000.0000.0001.837 (5)
Zr16c0.0000.0000.0000.163 (5)
Zr16d0.5000.5000.5001.837 (5)
Gd16d0.5000.5000.5000.163 (5)
O148f0.4110.1250.1256.00 (5)
O28a0.1250.1250.1251.00 (0)
Rp = 2.56, Rwp = 3.72, Rexp = 2.39, Rb = 3.67, Rf = 4.59, Chi2 = 2.41.
Table 2. The refined coordinates of atoms of 0.95Gd2Zr2O7–0.05MgO in a cubic unit cell (space group Fd-3m), with Gd and Mg existing in the A sublattice.
Table 2. The refined coordinates of atoms of 0.95Gd2Zr2O7–0.05MgO in a cubic unit cell (space group Fd-3m), with Gd and Mg existing in the A sublattice.
AtomWyckhoff Symbolx/ay/bz/cOccupancy
Gd16c0.0000.0000.0001.880 (5)
Zr16c0.0000.0000.0000.070 (5)
Mg16c0.0000.0000.0000.050
Zr16d0.5000.5000.5001.930 (5)
Gd16d0.5000.5000.5000.070 (5)
O148f0.4120.1250.1255.56 (2)
O28a0.1250.1250.1251.00 (0)
O38b0.3750.3750.3750.44 (2)
Rp = 2.46, Rwp = 3.54, Rb = 3.61, Rf = 4.19, Chi2 = 3.23.
Table 3. The relative density of Gd2Zr2O7 and (1-x)Gd2Zr2O7·xMgO samples.
Table 3. The relative density of Gd2Zr2O7 and (1-x)Gd2Zr2O7·xMgO samples.
Chemical Formulaρ, %
Gd2Zr2O796d24
0.95Gd2Zr2O7–0.05MgO96.40
0.93Gd2Zr2O7–0.07MgO96.83
0.90Gd2Zr2O7–0.10MgO97.27
Table 4. Photographs of the obtained ceramic samples of (1-x)Gd2Zr2O7·xMgO before and after exposure in LiCl-xLi2O melts at T = 650 °C.
Table 4. Photographs of the obtained ceramic samples of (1-x)Gd2Zr2O7·xMgO before and after exposure in LiCl-xLi2O melts at T = 650 °C.
x in
(1-x)Gd2Zr2O7·xMgO
Photo before the ExperimentPhoto after 48 h of Exposure in LiClPhoto after 48 h of Exposure in LiCl-2 wt.% Li2OPhoto after 48 h of Exposure in LiCl-4 wt.% Li2O
0.05Processes 11 01217 i001Processes 11 01217 i002Processes 11 01217 i003Processes 11 01217 i004
0.07Processes 11 01217 i005Processes 11 01217 i006Processes 11 01217 i007Processes 11 01217 i008
0.10Processes 11 01217 i009Processes 11 01217 i010Processes 11 01217 i011Processes 11 01217 i012
Table 5. Change in the mass loss of ceramic samples of (1-x)Gd2Zr2O7·xMgO before and after exposure in LiCl-xLi2O melts at T = 650 °C.
Table 5. Change in the mass loss of ceramic samples of (1-x)Gd2Zr2O7·xMgO before and after exposure in LiCl-xLi2O melts at T = 650 °C.
CompositionBeforeAfter
LiCl
After
LiCl-2 wt.% Li2O
After
LiCl-4 wt.% Li2O
%%%%
0.95Gd2Zr2O7-0.05MgO-+0.28−0.07−3.4
0.93Gd2Zr2O7-0.07MgO-−0.01+0.09−2.89
0.90Gd2Zr2O7-0.10MgO-−0.32−2.73−11.46
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

Anokhina, I.; Animitsa, I.; Erzhenkov, M.; Voronin, V.; Kadyrova, N.; Zaikov, Y. Electrical Properties and Chemical Resistance of the Composites (1-x)Gd2Zr2O7·xMgO in Li-Containing Chloride Melts. Processes 2023, 11, 1217. https://doi.org/10.3390/pr11041217

AMA Style

Anokhina I, Animitsa I, Erzhenkov M, Voronin V, Kadyrova N, Zaikov Y. Electrical Properties and Chemical Resistance of the Composites (1-x)Gd2Zr2O7·xMgO in Li-Containing Chloride Melts. Processes. 2023; 11(4):1217. https://doi.org/10.3390/pr11041217

Chicago/Turabian Style

Anokhina, Irina, Irina Animitsa, Maxim Erzhenkov, Vladimir Voronin, Nadezhda Kadyrova, and Yuri Zaikov. 2023. "Electrical Properties and Chemical Resistance of the Composites (1-x)Gd2Zr2O7·xMgO in Li-Containing Chloride Melts" Processes 11, no. 4: 1217. https://doi.org/10.3390/pr11041217

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop