Next Article in Journal
Modeling and Characterization of Surface Discharges in Insulating Material for Spacers: Electrode Shape, Discharge Mode, and Revision of the Creepage Concept
Next Article in Special Issue
Structure and Properties of Porous Ti3AlC2-Doped Al2O3 Composites Obtained by Slip Casting Method for Membrane Application
Previous Article in Journal
Ni Porous Preforms Compacted with Al2O3 Particles and Al Binding Agent
Previous Article in Special Issue
(Na, Zr) and (Ca, Zr) Phosphate-Molybdates and Phosphate-Tungstates: II–Radiation Test and Hydrolytic Stability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

(Na, Zr) and (Ca, Zr) Phosphate-Molybdates and Phosphate-Tungstates: I–Synthesis, Sintering and Characterization

1
Materials Science Department, Physical and Technical Research Institute, Lobachevsky State University of Nizhny Novgorod, 603022 Nizhny Novgorod, Russia
2
Faculty of Chemistry, University of Oviedo, 33006 Oviedo, Spain
3
Laboratory of Diagnostics of Radiation Defects in Solid State Nanostructure, Institute for Physics of Microstructure, Russian Academy of Science, 603950 Nizhniy Novgorod, Russia
4
Center Collective Use “Materials Science and Metallurgy”, National University of Science and Technology “MISIS”, 119991 Moscow, Russia
5
Laboratory “FIANIT”, Laser Materials and Technology Research Center, A.M. Prokhorov General Physics Institute of the Russian Academy of Sciences, 119991 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(3), 990; https://doi.org/10.3390/ma16030990
Submission received: 28 November 2022 / Revised: 29 December 2022 / Accepted: 17 January 2023 / Published: 20 January 2023

Abstract

:
Submicron-grade powders of Na1-xZr2(PO4)3-x(XO4)x compounds (hereafter referred to as NZP) and Ca1-xZr2(PO4)3-x(XO4)x compounds (hereafter, CZP), X = Mo, W (0 ≤ x ≤ 0.5) were obtained by sol-gel synthesis. The compounds obtained were studied by X-ray diffraction phase analysis and electron microscopy. An increase in the W or Mo contents was shown to result in an increase in the unit cell volume of the NZP and CZP crystal lattices and in a decrease in the coherent scattering region sizes. Thermal expansion behavior at high temperatures of synthesized NZP and CZP compounds has been investigated. The dependencies of the parameters a and c on the heating temperature, as well as the temperature dependence of the crystal lattice unit cell volume V in the range from the room temperature up to 800 °C, were obtained. The dependencies of the average thermal expansion coefficient (αav) and of the volume coefficient (β) on the W and Mo contents in the compositions of NZP and CZP compounds were studied. Ceramics Na1-xZr2(PO4)3-x(XO4)x with relatively high density (more than 97.5%) were produced by spark plasma sintering (SPS). The increase in the W or Mo contents in the ceramics leads to an increase in the relative density of NZP and to a decrease of the optimum sintering temperature. The mean grain size in the NZP ceramics decreases with increasing W or Mo contents. The study of strength characteristics has revealed that the hardness of the NZP ceramics is greater than 5 GPa, and that the minimum fracture toughness factor was 1 MPa·m1/2.

1. Introduction

In this paper, the inorganic compounds with the NaZr2(PO4)3 (NZP-type) structure are studied extensively in connection with the prospects of application for solving the problems of immobilization of the highly active components of radioactive waste (HLW) [1,2,3,4]. In particular, the Mo-containing fractions of the radioactive waste solidify as inclusions in glass, together with other nuclides, for example, the wedge-shaped glass ceramics in the borosilicate glasses [5,6,7]. When more than 10 wt.% of Mo is incorporated into a multi-component highly radioactive waste, Mo can form powellite—a mineral of Ca molybdate, CaMoO4. This affects the chemical resistance of the glasses, adversely affecting the solubility of molybdates [7]. A similar situation takes place with W.
Ceramics with an NZP-type structure can be quite effective at binding W and Mo into stable crystalline compounds where W and Mo can partially replace P [8,9,10,11,12,13,14,15,16]. Stable crystalline NZP-like matrix substances are most suitable for binding Mo and W. The inclusion of Mo in NZP-like ceramics leads to a decrease in the Mo leaching rate as compared with the glasses, glass ceramics, and Synroc materials containing individual phases of readily soluble molybdates [8]. The sintering of the crystalline compounds into ceramics increases their relative density, while preserving the phase composition. At the same time, the resulting ceramics have high strength characteristics [1,4,17,18,19,20].
The compounds with the NaZr2(PO4)3 structure are characterized by the crystal chemical formula (M1)VI(M2)VIII3[LVI2(XO4)3]n-,here M1 and M2 are the sites in the voids of the [LVI2(XO4)3]n- framework. The family of phosphates with the NZP structure includes many compounds, due to the possibility of isomorphic substitutions in various positions of the structure [21,22,23,24,25,26,27,28,29,30,31,32,33]. The framework of the structure is formed by multiply charged cations L with oxidation states 5+, 4+, 3+, or 2+ of small sizes and by anions XO4-. Most compounds of the NZP family contain P as an anion-forming element X. However, there are also compounds with the NZP-type structure, in which P is replaced by other anions. Some compounds are known where P is replaced by Si [17,24,25,26,27], by S [28,29], by V [30], by As [31], by Se [32], and by Mo [8,9,33].
The M sites can be occupied fully or partially, or remain vacant. The composition of NZP-type phosphates can include cations in the oxidation states from +1 to +4; mainly, low-charged and relatively large cations occur. The presence of four positions (M1, M2, L, and X) capable of incorporating cations of various sizes determines the prospects for the application of materials based on compounds with the NZP-type structure in various fields. The compounds of this structure are characterized by high ionic conductivity as well as by high corrosion, thermal, radiation, and chemical resistance, along with high catalytic activity [33,34].
The behavior of phosphates under heating and the values of thermal expansion coefficients depend on the charge, size, and electronegativity of their constituent ions. Due to the ability of the structure to accommodate various constituent elements, it becomes possible to develop various materials with desired thermal expansion coefficients [35]. In most cases, these compounds are characterized by expansion of the unit cell along the c crystallographic axis and by compression along the a and b axes when heated (anisotropic thermal expansion). In the case of a partial or complete substitution of the anions, the charge (n) of the framework changes. In this case, the cations in the Msites compensate for this charge, preserving the electroneutrality. Therefore, the cations in the M sites and the total occupancy of the sites affect the changes in the thermal expansion coefficients. Some of these compounds have small or ultra-low [down to (1–2)·10−6 deg−1] controlled thermal expansion coefficients [1,4,36,37,38,39,40,41,42,43]. Also, these compounds are stable under hydrothermal conditions at temperatures up to 400 °C and at durations of contact with water up to two years [1,4,18,44,45,46]. Such a unique hydrolytic resistance of the ceramics with the NZP-type structure increases interest in them as materials for solving the radiochemical problems of immobilization of radioactive waste [1,4,21].
At present, hot pressing or conventional sintering of preliminary pressed powders are applied for sintering the mineral-like ceramics most often [18,19,20,47]. The necessity of maintaining elevated sintering temperatures for long periods to provide a high density of the ceramics is the drawback of these methods. Spark plasma sintering (SPS) stands out from other methods of obtaining the ceramic samples. The method consists of high-speed heating (up to 2500 °C/min) of a powder material in vacuum by passing a series of short electric current pulses through a sample and a mold, while applying pressure [48,49,50,51,52]. This method is characterized by high shrinkage rates at reduced sintering temperatures. The sintering of the mineral-like ceramic samples up to high relative densities (about 95%) takes place in shorter times [27,51,52,53]. This ensures the reduction of the times of handling the highly active components of the radioactive waste as compared to hot pressing methods and conventional sintering of the powders pressed in advance. These advantages of SPS allow a significant expansion of applications in the development of ceramic matrices for the immobilization of high-level waste (HLW) and in the transmutation of minor actinides [4,27,51,52,53,54,55,56].
Ceramic materials obtained by SPS are characterized by a high relative density and enhanced physical and mechanical properties, which open up new possibilities for obtaining the ceramic materials for various purposes [4,27,48,49,50,51,52]. Some of these ceramics have a high radiation resistance [4,52,53,57] and high hydrolytic stability [4,56,58,59].
This work is devoted to the synthesis and investigation of properties of the NZP-framework orthophosphates with partial substitution of the P ions by Mo and W and various cations in the structure voids. These ceramics may potentially be applied to the immobilization of the Mo- and W-containing fractions of radioactive waste.

2. Materials and Methods

Solid solutions of the Na1-xZr2(PO4)3-x(XO4)x and Ca1-xZr2(PO4)3-x(XO4)x types, where X = Mo, W, and x = 0.1, 0.2, 0.3, 0.4, 0.5, were chosen for investigation. The compounds were synthesized by the sol-gel method. A sample of Zr oxychloride was dissolved in a mixture of solutions of ammonium molybdate or tungstate and Na or Ca nitrate. A 1 M solution of ammonium dihydrogen phosphate was added dropwise to the resulting solution in the course of continuous stirring. As a result, gel-like precipitates were formed. The reagents were mixed in the stoichiometric proportions. For more complete precipitation, a salting-out agent (ethyl alcohol) was added. After brief stirring, the gel was dried at 90 °C for 1 day. The gel-like sediment was dried in a drying cabinet and then was dispersed in an agate mortar, heated up to 600 °C, dispersed again, and heated up to 800 °C. The resulting powders were investigated by X-ray diffraction (XRD) between the stages of annealing.
The ceramics were sintered from the obtained powders by SPS using Dr. Sinter™ model SPS-625 (SPS SYNTEX®, Kanagawa, Japan). Prior to sintering, the powders were ground in an agate mortar for 5 min, and pressed into a graphite mold at a pressure of 70 MPa. No binders were used for the preparation of the initial powder green body. The binderless powders were placed in a graphite mold with an inner diameter of 12.8 mm and an outer diameter of 30 mm, and heated up by passing millisecond pulses of high-power direct electric current (up to 5 kA). Sintering was performed in a vacuum. The effective temperature (T1) was measured with a Chino IR-AHS2 pyrometer (Chino Corporation, Tokyo, Japan) focused on the graphite mold surface. Based on previous studies, and on the comparison of the infrared pyrometer readings (T1 in Figure 1) to those of a thermocouple attached to the ceramic specimen surface, the values of T1 were used to calculate the actual specimen temperature (T) using the following empirical relation: T = A·T1 − B, where A and B are empirical constants which depend on the heating rate Vh (Figure 1 The method of comparing temperatures T and T1 is described in [60]. To improve the contact of the powder sample with the graphite mold and to compensate for the difference in the thermal expansion coefficients of graphite and the ceramics, a graphite foil was placed inside the graphite mold. A two-stage heating regime was used—stage A: heating up to 570–580 °C with the heating rate Vh1 = 100 °C/min, and stage B: heating up to the sintering temperature Ts with the heating rate Vh2 = 50 °C/min. Holding at the sintering temperature Ts was absent (ts = 0 s). The average uniaxial stress applied was ~65 MPa. In order to avoid cracking of the ceramic specimen, multistage pressure regulation was applied during sintering: (stage I) a gradual increasing of the uniaxial stress P from 35–38 MPa up to 64–65 MPa during the rapid heating stage up to 570–580 °C, (stage II) holding at 64–66 MPa at 580–840 °C, (stage III) reducing the pressure to 60 MPa at 900 °C, and (stage IV) gradually increasing the pressure again to ~65 MPa to the moment of the end of shrinkage. The typical scheme of the stress and temperature regulation during SPS is presented in Figure 1. The samples were cooled down together with the setup (stage C, Figure 1). As a result of sintering, dense ceramic specimens without macrodefects were obtained.
The temperature curves of the effective shrinkage (Leff) were measured during sintering using the built-in dilatometer of the SPS-625. To account for thermal expansion of the mold (L0), a separate series of experiments on heating up an empty mold was carried out. True shrinkage was calculated as L = Leff – L0 (Figure 2). Using the dependencies L(T), the temperature dependencies of the shrinkage rate were calculated: S = ΔL/Δt.
The optimal sintering temperature Ts for each ceramic was selected on the basis of analysis of the dependencies Leff(T). The samples were heated up to the temperature corresponding to the end of the stage of intensive powder shrinkage. Due to the contribution of the thermal expansion L0(T) on the Leff(T) curve, stage III arose in the temperature dependence of true shrinkage L(T), within which true shrinkage of the ceramics almost didn’t change (Figure 2). The change in temperature during stage III was ~50 °C, corresponding to the time interval of ~1 min at the heating rate Vh2 = 50 °C/min. In the course of heating, we tried to minimize the duration of stage III, since intensive grain growth takes place in the ceramics at high temperatures (T > Ts). The coarse-grain ceramics have very low resistance against thermal shock, and the samples are often destroyed during cooling down. Additionally, the sintering at high temperatures can lead to partial dissociation of the elements (W, Mo) from the ceramic surfaces and to more intensive saturation of the ceramic sample surfaces with the mold (see [50,61,62,63,64,65,66,67]).
The XRD phase analysis of the ceramics was performed using Bruker® D8 Discover™ X-ray diffractometer (Bruker Corp., Billerica, MA, USA) in symmetric Bragg-Brentano geometry. An X-ray tube with a Cu cathode (CuKα radiation) was used. The goniometer radius was 30 cm. All experiments were performed in the same conditions. The specimens, comprising cylindrical tablets of 12 mm diameter and 3 mm height, were placed on the goniometer table. At the beginning of each experiment, adjustment of the specimen height by the beam splitting method was performed with 0.1 mm slits both on the primary beam and in front of the detector. When acquiring the XRD curves, the primary beam size was limited in the equatorial plane by a 0.6 mm slit, and in the axial plane by a 12 mm slit. The XRD curves were acquired in the θ/2θ-scan mode in the angle range from 10° to 70° in 2θ. The step in angle was 0.1°. LynxEYE linear position-sensitive detector with 192 independent acquisition channels and a 2° angle aperture in 2θ was used. The dwell time was 2 s, which was equivalent to the effective dwell time about 40 s due to the use of the position-sensitive detector. The thermal expansion in the temperature range 25–800 °C was investigated using Panalytical® X’Pert Pro™ high-temperature X-ray diffractometer (Malven PANAlytical B.V., Almelo, The Netherlands) with Anton Paar® HTK–1200N high-temperature chamber (Anton Paar GmnH, Graz, Austria) with a step of 100 °C. The unit cell parameters were calculated with the Rietveld method using Topas 4.2 software.
The microstructure of the powders and ceramics was investigated using Jeol® JSM-6490 scanning electron microscope (SEM, Jeol Ltd., Tokyo, Japan) and Jeol JEM-2100 transmission electron microscope (TEM, Jeol Ltd., Tokyo, Japan). The mean particle size (DP), the size of agglomerates (DA), and grain size (d) were calculated by the section method using GoodGrains 2.0 software (UNN, Nizhny Novgorod, Russia). To determine the mean particle sizes, investigations of the sizes of all particles of the powders presented in the images obtained by TEM were conducted. When determining the mean grain sizes, we measured 100 or more grains presented in the images obtained by SEM. When determining DP and d, three photographs (images) or more were analyzed. The mean sizes were calculated with GoodGrains 2.0 software in the automated mode, on the base of analysis of the histograms of the size distributions of the particles/grains. The uncertainties of determining DP and d were calculated as the standard deviations.
The density of the obtained ceramics was measured by hydrostatic weighing in distilled water using Sartorius® CPA 225D balance (Sartorius AG, Göttingen, Germany). The uncertainty of the density measurement was ±0.001 g/cm3. The theoretical density of the ceramics (ρth) was calculated on the basis of the XRD investigations. The microhardness (Hv) of the ceramics was measured using Duramin® Struers-5 hardness tester (Struers LLC, Cleveland, OH, USA). The load was 200 g. The minimum fracture toughness factor (KIC) was calculated according to the Palmquist method from the length of the largest radial crack forming by the Vickers pyramid during the indentation of the ceramics:
KIC = 0.016(P/c3/2)(E/Hv)1/2,
where P is the indenter load [g], c is the mean distance from the imprint center to the end of the crack, and E is the elastic modulus of the material [GPa]. When calculating the magnitude of KIC, we used the lengths of 10 or more of the longest cracks from 10 different indents formed on the ceramic sample surfaces when loading by Vickers pyramid.

3. Results and Discussion

3.1. Synthesis and Characterization of the Powders

The powder samples were light-colored. According to the SEM results, the submicron-grained powders were agglomerated; there were some large agglomerations up to 10–20 μm in size in the Ca1-xZr2(PO4)3-x(XO4)x powders (denoted as “CZP” in work) (Figure 3a) and ~50 μm in size in the Na1-xZr2(PO4)3-x(XO4)x powders (denoted as “NZP”) (Figure 3c,e) after grinding in the agate mortar. No considerable effect of the W and Mo contents on the granulometric composition and morphology of the powders was observed.
Figure 4 and Figure 5 present TEM images of the Ca1-xZr2(PO4)3-x(MoO4)x powders with different Mo contents (Figure 4) and of Na0.5Zr2(PO4)2.5(XO4)0.5 powders, X = Mo, W (Figure 5). The Ca1-xZr2(PO4)3-x(MoO4)x powders were large-grained and agglomerated strongly. The individual particle sizes ranged from ~50–100 nm to ~300 nm and didn’t depend on the Mo content (Figure 4). The Na0.5Zr2(PO4)2.5(XO4)0.5 powders were more finely dispersed and agglomerated. The particle sizes ranged from ~20–30 nm to ~50–100 nm (Figure 5). The agglomerate sizes were ~0.2–0.3 μm. All powders had crystalline structure (Figure 4b,d,f and Figure 5b,d,e,f).
According to the XRD phase analysis data (Figure 6 and Figure 7), the compounds under study crystallized in the NaZr2(PO4)3 type structure, hexagonal syngony, space group R 3 ¯ c (analog of NaZr2(PO4)3 [37]) for the Na-containing samples and space group R 3 ¯ (analog of Ca0.5NaZr2(PO4)3 [68]) for the Ca-containing ones. The XRD patterns of the samples with Ca at x > 0.2 for phosphate–molybdates and x > 0.3 for phosphate–tungstates contained a significant number of auxiliary (secondary) phase reflections. For this reason, these samples were not studied further.
The particle (coherent scattering region, CSR) sizes DXRD were estimated using the Scherer equation:
DXRD = K·λ/β·cosθmax,
where K = 0.9 is the Scherer constant, λ is the wavelength of the X-ray radiation, β is the full width at half maximum of the XRD peak (in radians), and θmax is the diffraction angle corresponding to the (116) XRD peak maximum. The analysis of the XRD results demonstrated that the particle sizes DXRD of the Na1-xZr2(PO4)3-x(MoO4)x powders decreased monotonously from 61 nm down to 19 nm with increasing Mo content from 0.1 up to 0.5 and from 44 nm down to 23 nm with increasing W content from 0.1 up to 0.5 (Figure 8). No considerable effects of W and Mo content on the DXRD of the Ca1-xZr2(PO4)3-x(XO4)x powders were observed; the mean particle sizes were ~60 nm for the W-containing powders and ~47 nm for the Ca1-xZr2(PO4)3-x(MoO4)x ones.
From the analysis of the XRD, TEM, and SEM results comes the conclusion that partial sintering of several nanoparticles into submicron ones takes place in the course of synthesis, at the last stage of high-temperature annealing (800 °C, 20 h). This conclusion was confirmed by comparing the results of investigations of the mean particle sizes by XRD, TEM, and SEM. The analysis of the XRD results shows that the mean CSR sizes in the NZP and CZP compounds were several tens of nanometers (DXRD < 60 nm). The results of TEM and SEM show the particles to be agglomerated strongly; the agglomerates of submicron and micron sizes consist of separate nanoparticles (Figure 2, Figure 3 and Figure 4).
The unit cell parameters for the solid solutions were determined from the XRD data (the results are presented in Table 1).
The composition dependences of the unit cell parameters for the phosphate–molybdates and phosphate–tungstates are presented in Figure 9. Substitution of phosphate anions (PO4)3− (R (P5+) = 0.17 Å) by larger molybdate anions (MoO4)2− (R (Mo6+) = 0.41 Å) and tungstate ions (WO4)2− (R(W6+) = 0.44 Å) leads to an increase in the unit cell parameters. One can see from Figure 9 that increasing the W content in the Na1-xZr2(PO4)3-x(XO4)x compounds results in a greater increase in the lattice parameter c and the unit cell volume V than increasing the Mo content. No essential differences in the values of the lattice parameter a for the phosphate–molybdates and phosphate–tungstates Na1-xZr2(PO4)3-x(XO4)x were observed. For the Ca–phosphate–tungstates Ca1-xZr2(PO4)3-x(WO4)x, no dependencies of the unit cell parameters a and c, nor the unit cell volume V, on the W content were found. In the case of Ca1-xZr2(PO4)3-x(MoO4)x, the increase in the Mo content up to x = 0.2 resulted in a linear increase in all unit cell parameters and, hence, in a linear increase in the unit cell volume V.
To study the behavior of the obtained compounds upon heating, the XRD patterns of the samples were recorded at elevated temperatures (RT–800 °C). The XRD data were used to calculate the values of the unit cell parameters at different temperatures. The temperature dependencies of the cell parameters are shown in Figure 10 and Figure 11. One can see from Figure 10 a minor decrease (~0.02 Å) in the unit cell parameter a and a considerable increase in the parameter c (~0.2–0.4 Å) with increasing temperature (from 20 up to 800 °C) for Na1-xZr2(PO4)3-x(XO4)x. This leads to an increase in the unit cell volume V when heating. For Ca1-xZr2(PO4)3-x(XO4)x, a greater increase in the parameter c was observed with a similar increase in temperature (Figure 11), which resulted in a smaller increase in the unit cell volume of Ca1-xZr2(PO4)3-x(XO4)x as compared to Na1-xZr2(PO4)3-x(XO4)x.
The above dependencies were used to calculate the values of the axial (αa and αc), average (αav), and volume (β) thermal expansion coefficients (CTEs), as well as the thermal expansion anisotropy (Δα) for the phosphate–molybdates and phosphate–tungstates under study (Table 2, Figure 12 and Figure 13). The analysis of these data shows the decreasing of αa and αc with increasing W and Mo contents in Na1-xZr2(PO4)3-x(XO4)x. At the same time, the mean value of αav changes slowly. The volume CTE (β) decreases slowly as the thermal expansion anisotropy (Δα) increases. In general, a similar character of the CTE dependencies on the composition was observed for the Ca1-xZr2(PO4)3-x(MoO4)x compounds with the increase I n the Mo content (Figure 12). For the W-containing phosphates Ca1-xZr2(PO4)3-x(WO4)x, no considerable effect of the W content on all thermal expansion parameters (αa, αc, αav, β, and Δα) was observed.
One can note a trend for the thermal expansion parameters to approach zero with decreasing population of the extraframe positions in the structure of the phosphate–molybdates and phosphate–tungstates under study. Therefore, novel ceramic materials with zero CTE and high thermal shock resistance can potentially be developed on the basis of the phosphates with predefined W and/or Mo contents studied in the present work.

3.2. Sintering and Characterization of the Ceramics

The ceramic samples with high relative densities were obtained from the Na-containing compounds by SPS. The ceramic samples have no visible macro- nor microcracks. Examples of temperature curves L(T) and S(T) are shown in Figure 14. The shapes of the curves L(T) are typical for SPS of the nano- and submicron NZP-type powders [27,51,53].
The analysis of the temperature curves of the shrinkage L(T) for the phosphate–molybdates (Figure 14a,c) shows an increase in the maximum shrinkage Lmax and of the shrinkage rate Smax with increasing Mo content from x = 0.1 up to 0.3. At the same time, an increase in the optimal sintering temperature was observed, which is manifested as the shift of the curves L(T) and S(T) towards higher heating temperatures. With a further increase in the Mo content up to x = 0.5, a shift of the curves L(T) and S(T) towards lower sintering temperatures was observed. In our opinion, such a behavior is due, first of all, to the appearance of some auxiliary phases during sintering, and to the increasing concentration of the auxiliary phases with increasing Mo concentration (see above).
In the Na1-xZr2(PO4)3-x(WO4)x phosphates, a minor increase in the shrinkage and a decrease in the optimal sintering temperature corresponding to the completion of the intensive shrinkage stage with increasing W content were observed (Figure 14b). At the same time, an increase in the maximum shrinkage rate of nanopowders Smax of more than three times, and a minor shift of the temperatures, during which the maximum shrinkage rates Smax were manifested towards higher heating temperatures, were observed (Figure 14d).
The average full SPS process times (tsps), including the heating times, were 13 min for the phosphate–molybdates Na1-xZr2(PO4)3-x(MoO4)x and 16 min for the phosphate–tungstates Na1-xZr2(PO4)3-x(WO4)x. The characteristics of the sintering process (the heating rates Vh, the sintering temperatures Ts, and the full SPS process time tsps), the relative densities achieved, the microhardness values, and the minimal fracture toughness factors for the ceramics obtained are presented in Table 3. One can see from Table 3 that the increasing of the W and Mo contents resulted in an increasing of the relative density of the Na1-xZr2(PO4)3-x(XO4)x ceramics. Densities close to the theoretical one were ensured for almost all ceramics. In the ceramics with increased W and Mo contents, the relative densities exceeded 100% slightly. In our opinion, this result originates from the auxiliary (secondary) phases contained in the powders synthesized (see the XRD phase analysis results above). The results of XRD phase analysis of the ceramics also show the secondary phases to be present in the ceramics (Figure 15). In the W-containing ceramics, the presence of Zr2(WO4)(PO4)2 secondary phase was observed (Figure 15a), in the Mo-containing ones CaCO3 and Al3O0.34Zr5, secondary phases were found (Figure 15b).
One can see from Table 3 that in the NZP ceramics with high Mo and W contents, one can provide a high density at reduced sintering temperatures. We suggest the that an increase in Mo and W contents will result in a decrease in the sintering activation energy for the Na1-xZr2(PO4)3-x(XO4)x ceramics, X = Mo, W (0 ≤ x ≤ 0.5). This allows sintering the ceramics with high values of x up to high densities at lower temperatures. The decrease in the sintering activation energy was manifested indirectly also in a decrease in the temperature at the beginning of the intensive shrinkage stage (Stage II) in Figure 14 at low heating temperatures.
Figure 16 and Figure 17 present the SEM images of the microstructure of the sintered phosphate–molybdates Na1-xZr2(PO4)3-x(MoO4)x (Figure 16) and phosphate–tungstates Na1-xZr2(PO4)3-x(WO4)x (Figure 17) with various Mo and W contents. One can see in Figure 16 a decrease in the mean grain size (Figure 16b,d) and an increase in the microstructure uniformity of the ceramic phosphate–molybdates (Figure 16a,c) with increasing Mo content. Note that there were some areas with abnormally large grains in the sintered ceramics with low Mo contents (x = 0.1). The sizes of these grains were ~5–10 μm (Figure 16a), greater than the mean grain size of the ceramic matrix (Figure 16b) by an order of magnitude. In the ceramics with increased Mo contents (x = 0.5), the areas with abnormally large grains were almost absent (Figure 16c), whereas the mean grain sizes in the sintered ceramic phosphate-molybdates were close to 100–200 nm (Figure 16d).
A similar effect has been observed when analyzing the results of SEM investigations of the ceramic phosphate–tungstates with different W contents. As one can see in Figure 17, an increase in the W content resulted in a decrease in the mean grain sizes in the ceramics down to ~100–150 nm. There are few areas with abnormally large grains, the sizes of which is not greater than 2–5 μm in the ceramics with x = 0.5 (Figure 17d).
The XRD phase analysis evidenced the phase composition of the ceramics not to change after sintering (Figure 15). A minor increase in the intensities and decrease in the widths of the XRD peaks have been observed that is typical for sintering the nano- and submicron powders, which is associated with the grain growth.
As one can see from Table 3, the increase in the Mo and W contents resulted in a minor increase in the microhardness of the ceramics up to 5.3–5.6 GPa. At the same time, the minimum fracture toughness factor almost did not change and exceeded 1 MPa·m1/2 as a rule, which is a typical value for the mineral-like ceramics with the NZP structure [4,56].

4. Conclusions

1.
The submicron-grade powders of the Na1-xZr2(PO4)3-x(XO4)x and Ca1-xZr2(PO4)3-x(XO4)x compounds, X = Mo, W (0 ≤ x ≤ 0.5) were obtained by sol-gel synthesis. The powders were agglomerated; large agglomerates of several microns in sizes consisted of individual nano- and submicron-sized particles. The sizes of individual particles of the Ca1-xZr2(PO4)3-x(XO4)x compounds varied from 50–100 to 300 nm; the ones for the Na1-xZr2(PO4)3-x(XO4)x compounds varied from ~20–30 nm to ~50–100 nm.
2.
It follows from the analysis of the results obtained that the investigated compounds crystallize in the NZP-type structure, hexagonal syngony, R 3 ¯ c space group for the Na-containing phosphate–molybdates and phosphate–tungstates and R 3 ¯ for the Ca-containing ones. The substitution of the phosphate anions (PO4)3− (R(P5+) = 0.17 Å) by a larger molybdate (MoO4)2− (R (Mo6+) = 0.41 Å) and tungstate (WO4)2− (R(W6+) = 0.44 Å) ions leads to an increasing of the unit cell parameters. With a decrease in the population of the extraframe positions in the phosphate–molybdates and phosphate–tungstates investigated, the thermal expansion parameters tend to approach zero.
3.
The dependencies of the crystal lattice constants a and c as well as of the unit cell volume V on the heating temperature were studied by high-temperature X-ray diffraction. The dependencies of the mean thermal expansion coefficient (TEC) (αav) and of the volume coefficient (β) on the W and Mo contents in NZP and CZP compounds were investigated. The mean values of TEC (αav) for the Mo-containing NZP and CZP compounds are (2.97–4.41)·10−6 °C−1 and (1.45–1.55)·10−6 °C−1, respectively. The mean values of αav for the W-containing NZP and CZP compounds are (1.41–3.33)·10−6 °C−1 and (1.70–2.44)·10−6 °C−1, respectively. The increase in the tungsten content leads to a decrease in αav for the compounds with the NZP structure. The values of the volume TEC (β) and of the TEC anisotropy (Δα) for the NZP compounds decrease with increasing W content. The effects of Mo and W on the increase in β and Δα for the CZP compounds are insignificant.
4.
The bulk of the NZP ceramic specimens obtained by SPS were characterized by high relative densities (ρrel> 97.5%). The full time of the SPS process, including the heating and cooling-down times, is ~11.5–16.5 min for the Mo-containing ceramics and ~15.1–18.1 min for the W-containing ones. The values of microhardness (Hv) of the ceramics ranged from 3.6 to 5.6 GPa, and the minimal fracture toughness factor (KIC) ranged from 0.9 to 1.6 MPa m1/2. The ceramics with the NZP structure have fine-grained microstructure; abnormal grain growth was observed in the ceramics with increased W contents. In the NZP ceramics with high Mo and W contents, one can provide high density at reduced sintering temperature.

Author Contributions

Conceptualization, A.I.O. and A.V.N.; methodology, A.I.O.; validation, A.I.O., A.V.N. and V.N.C.; formal analysis, A.I.O. and A.V.N.; investigation, M.E.K., D.O.S., S.A.K., M.S.B., S.G.-G., A.A.M., P.A.Y., A.A.N. and N.Y.T.; resources, A.I.O. and V.N.C.; data curation, A.I.O. and A.V.N.; writing—original draft preparation, A.I.O., A.V.N. and V.N.C.; writing—review and editing, A.I.O., A.V.N. and V.N.C.; visualization, A.V.N.; supervision, A.I.O.; project administration, A.I.O.; funding acquisition, A.I.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation (Grant No. 21-13-00308). The TEM study of the powders was carried out on the equipment of the Center for Collective Use “Materials Science and Metallurgy” (National University of Science and Technology “MISIS”, Moscow, Russia) with the financial support of the Ministry of Science and Higher Education of the Russian Federation (Grant No. 075-15-2021-696). The XRD investigations of the specimens were carried out in the Laboratory of Diagnostics of Radiation Defects in Solid State Nanostructures at the Institute for Physics of Microstructures RAS (IPM RAS, Nizhny Novgorod, Russia) with the financial support of the Ministry of Science and Higher Education of the Russian Federation (Grant No. 0030-2021-0030).

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. Orlova, A.I.; Ojovan, M.I. Ceramic mineral waste-forms for nuclear waste immobilization. Materials 2019, 12, 2638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bohre, A.; Avasthi, K.; Pet’kov, V.I. Vitreous and crystalline phosphate high level waste matrices: Present status and future challenges. J. Ind. Eng. Chem. 2017, 50, 1–14. [Google Scholar] [CrossRef]
  3. Ewing, R.; Wang, L. Phosphates as Nuclear Waste forms. Review in Mineralogy and Geochemistry. Phosphates Geochem. Geobiol. Mater. Importance 2002, 48, 673–689. [Google Scholar] [CrossRef]
  4. Orlova, A.I. Crystalline phosphates for HLW immobilization—Composition, structure, properties and production of ceramics. Spark Plasma Sintering as a promising sintering technology. J. Nucl. Mater. 2022, 559, 153407. [Google Scholar] [CrossRef]
  5. Hayward, P.J.; Vance, E.R.; Cann, C.D.; George, I.M. Influence of heat treatment and redox conditions on crystallization of titanosilicate glass-ceramics. In Nuclear Waste Management 2 (Advances in Ceramics; V.20); Clark, D.E., White, W.B., Machiels, A.J., Eds.; American Ceramic Society: Westerville, OH, USA, 1986; pp. 215–222. [Google Scholar]
  6. Hayward, P.J.; Vance, E.R.; Cann, C.D. Crystallization of titanosilicate glasses for nuclear waste immobilization. J. Am. Ceram. Soc. 1989, 72, 579–586. [Google Scholar] [CrossRef]
  7. Sobolev, I.A.; Ojovan, M.I.; Shcherbatova, T.D.; Batyukhnova, O.G. Glass for Radioactive Waste; Energoatomizdat: Moscow, Russia, 1999; p. 238. (In Russian) [Google Scholar]
  8. Pet’kov, V.I.; Sukhanov, M.V.; Kurazhkovskaya, V.S. Molybdenum fixation in crystalline NZP matrices. Radiochemistry 2003, 45, 620–625. [Google Scholar] [CrossRef]
  9. Chourasiaa, R.; Shrivastavaa, O.P.; Wattal, P.K. Synthesis, characterization and structure refinement of sodium zirconium molibdato-phosphate: Na0.9Zr2Mo0.1P2.9O12 (MoNZP). J. Alloys Compd. 2009, 473, 579–583. [Google Scholar] [CrossRef]
  10. Bohre, A.; Shrivastava, O.P.; Avasthi, K. Solid state synthesis and structural refinement of polycrystalline phases: Ca1-2xZr4M2xP6-2xO24 (M = Mo, x = 0.1 and 0.3). Arab. J. Chem. 2016, 9, 736–744. [Google Scholar] [CrossRef] [Green Version]
  11. Kumar, S.P.; Gopal, B. Immobilization of “Mo6+” in monazite lattice: Synthesis and characterization of new phosphomolybdates, La1-xCaxP1-yMoyO4, where x = y = 0.1–0.9. J. Am. Ceram. Soc. 2011, 94, 1008–1013. [Google Scholar] [CrossRef]
  12. Daub, M.; Lehner, A.J.; Höppe, H.A. Synthesis, crystal structure and optical properties of Na2RE(PO4)(WO)4 (RE = Y,Tb-Lu). Dalton Trans. 2012, 41, 12121–12128. [Google Scholar] [CrossRef]
  13. Sukhanov, M.V.; Pet’kov, V.I.; Kurazhkovskaya, V.S.; Eremin, N.N.; Urusov, V.S. Computer-assisted structure simulation, synthesis, and phase formation of molybdophosphates A1-xZr2(PO4)3-x(MoO4)x (A is an alkali metal). Russ. J. Inorg. Chem. 2006, 51, 706–711. [Google Scholar] [CrossRef]
  14. Bennouna, L.; Arsalane, S.; Brochu, R.; Lee, M.R.; Chassaing, J.; Quarton, M. Spécificités des ions NbIV and MoIV dans les monophosphates de type Nasicon. J. Solid State Chem. 1995, 114, 224–229. [Google Scholar] [CrossRef]
  15. Buzlukov, A.L.; Fedorov, D.S.; Serdtsev, A.V.; Kotova, I.Y.; Tyutyunnik, A.P.; Korona, D.V.; Baklanova, Y.V.; Ogloblichev, V.V.; Kozhevnikova, N.M.; Denisova, T.A.; et al. Ion mobility in triple sodium molybdates and tungstates with a NASICON structure. J. Exp. Theor. Phys. 2022, 134, 42–50. [Google Scholar] [CrossRef]
  16. Vereshchagina, T.A.; Fomenko, E.V.; Vasilieva, N.G.; Solovyov, L.A.; Vereshchagin, S.N.; Bazarova, Z.G.; Anshits, A.G. A novel layered zirconium molybdate as a precursor to a ceramic zirconomolybdate host for lanthanide bearing radioactive waste. J. Mater. Chem. 2011, 21, 12001–12007. [Google Scholar] [CrossRef]
  17. Oikonomou, P.; Dedeloudis, C.; Stournaras, C.J.; Ftikos, C. [NZP]: A new family of ceramics with low thermal expansion and tunable properties. J. Eur. Ceram. Soc. 2007, 27, 1253–1258. [Google Scholar] [CrossRef]
  18. Orlova, A.I.; Zyryanov, V.N.; Kotelnikov, A.R.; Demarin, V.T.; Rakitina, E.V. Ceramic phosphate matrices for high-level waste. Behavior under hydrothermal conditions. Radiochemistry 1993, 35, 120–126. (In Russian) [Google Scholar]
  19. Arinicheva, Y.; Clavier, N.; Neumeier, S.; Podor, R.; Bukaemskiy, A.; Klinkenberg, M.; Roth, G.; Dacheux, N.; Bosbach, D. Effect of powder morphology on sintering kinetics, microstructure and mechanical properties of monazite ceramics. J. Eur. Ceram. Soc. 2018, 38, 227–234. [Google Scholar] [CrossRef]
  20. Dey, A.; Das Gupta, A.; Basu, D.; Ambashta, R.D.; Wattal, P.K.; Kumar, S.; Sen, D.; Mazumder, S. A comparative study of conventionally sintered, microwave sintered and hot isostatic press sintered NZP and CZP structures interacted with fluoride. Ceram. Int. 2013, 39, 9351–9359. [Google Scholar] [CrossRef]
  21. Orlova, A.I. Isomorphism in NZP-like phosphates and radiochemical problems. Radiochemistry 2002, 44, 385–403. [Google Scholar] [CrossRef]
  22. Volkov, Y.F.; Tomilin, S.V.; Orlova, A.I.; Lizin, A.A.; Spiryakov, V.I.; Lukinykh, A.N. Rhombohedral of actinide phosphates AIMIV2(P04)3 (MIV = U, Np, Pu; AI = Na, K, Rb). Radiochemistry 2003, 45, 319–328. [Google Scholar] [CrossRef]
  23. Orlova, A.I.; Koryttseva, A.K. Phosphates of pentavalent elements: Structure and properties. Crystallogr. Rep. 2004, 49, 811–819. [Google Scholar] [CrossRef]
  24. Alamo, J.; Roy, R. Ultralow-expansion ceramics in the system Na2O-ZrO2-P2O5-SiO2. J. Am. Ceram. Soc. 1984, 67, 78–80. [Google Scholar] [CrossRef]
  25. Tantri, S.P.; Ushadevi, S.; Ramasesha, S.K. High temperature X-ray studies on barium and strontium zirconium phosphate based low thermal expansion materials. Mater. Res. Bull. 2002, 37, 1141–1147. [Google Scholar] [CrossRef]
  26. Orlova, A.; Kanunov, A.E.; Samoilov, S.G.; Kazakova, A.Y.; Kazantsev, G.N. Study of calcium-containing orthophosphates of the structural type NaZr2(PO4)3 by high-temperature X-ray diffraction methods. Crystallogr. Rep. 2013, 58, 185–190. [Google Scholar] [CrossRef]
  27. Savinykh, D.O.; Khainakov, S.A.; Boldin, M.S.; Orlova, A.I.; Aleksandrov, A.A.; Lantsev, E.A.; Sakharov, N.V.; Murashov, A.A.; Garcia-Granda, S.; Nokhrin, A.V.; et al. Preparation of NZP-type Ca0.75+0.5xZr1.5Fe0.5(PO4)3-x(SiO4)x powders and ceramics, thermal expansion behavior. Inorg. Mater. 2018, 54, 1267–1273. [Google Scholar] [CrossRef]
  28. Alamo, J.; Roy, R. Zirconium phospho-sulfates with NaZr2(PO4)3-type structure. J. Solid State Chem. 1984, 51, 270–273. [Google Scholar] [CrossRef]
  29. Savinykh, D.O.; Khainakov, S.A.; Orlova, A.I.; Garcia-Granda, S. New phosphate-sulfates with NZP Structure. Russ. J. Inorg. Chem. 2018, 63, 685–694. [Google Scholar] [CrossRef]
  30. Pet’Kov, V.I.; Sukhanov, M.V.; Shipilov, A.S.; Kurazhkovskaya, V.S.; Borovikova, E.Y.; Sakharov, N.V.; Ermilova, M.M.; Orekhova, N.V. Synthesis and structure of vanadate phosphates of zirconium and alkaline metals. Russ. J. Inorg. Chem. 2013, 58, 1139–1145. [Google Scholar] [CrossRef]
  31. Sukhanov, M.V.; Pet’Kov, V.I.; Firsov, D.V.; Kurazhkovskaya, V.S.; Borovikova, E.Y. Synthesis, structure and thermal expansion of sodium zirconium arsenate phosphates. Russ. J. Inorg. Chem. 2011, 56, 1351–1357. [Google Scholar] [CrossRef]
  32. Slater, P.R.; Greaves, C. Synthesis and conductivities of sulfate/selenite phases related to Nasicon: NaxM’(II)xM’’(III)2-x(SO4)3-y(SeO4)y. J. Solid State Chem. 1993, 107, 12–18. [Google Scholar] [CrossRef]
  33. Petkov, V.I.; Orlova, A.I.; Egorkova, O.V. On the existence of phases with a structure of NaZr2(PO4)3 in series of binary orthophosphates with different alkaline element to zirconium ratios. J. Struct. Chem. 1996, 37, 933–940. [Google Scholar] [CrossRef]
  34. Il’In, A.B.; Novikova, S.; Sukhanov, M.; Ermilova, M.M.; Orekhova, N.V.; Yaroslavtsev, A.B. Catalytic activity of NASICON-type phosphates for ethanol dehydration and dehydrogenation. Inorg. Mater. 2012, 48, 397–401. [Google Scholar] [CrossRef]
  35. Pet’kov, V.I.; Orlova, A.I. Crystal-chemical approach to predicting the thermal expansion of compounds in the NZP family. Inorg. Mater. 2003, 39, 1013–1023. [Google Scholar] [CrossRef]
  36. Lenain, G.E.; McKinstry, H.A.; Limaye, S.Y.; Woodward, A. Low thermal expansion of alkali-zirconium phosphates. Mater. Res. Bull. 1984, 19, 1451–1456. [Google Scholar] [CrossRef]
  37. Limaye, S.Y.; Agrawal, D.K.; McKinstry, H.A. Synthesis and Thermal Expansion of MZr4P6O24 (M = Mg, Ca, Sr, Ba). J. Am. Ceram. Soc. 1987, 70, 232–236. [Google Scholar] [CrossRef]
  38. Volgutov, V.Y.; Orlova, A.I. Thermal expansion of phosphates with the NaZr2(PO4)3 structure containing lanthanides and zirconium: R0.33Zr2(PO4)3 (R = Nd, Eu, Er) and Er0.33(1-x)Zr0.25xZr2(PO4)3. Crystallogr. Rep. 2015, 60, 721–728. [Google Scholar] [CrossRef]
  39. Carrasco, M.P.; Guillem, M.C.; Alarmo, J. Preparation and structural study of sodium germanium phosphate-sodium titanium phosphate solid solutions II. Evolution of thermal expansion with composition. Mater. Res. Bull. 1994, 29, 817–826. [Google Scholar] [CrossRef]
  40. Oota, T.; Yamai, I. Thermal Expansion Behavior of NaZr2(PO4)3Type Compounds. J. Am. Ceram. Soc. 1986, 69, 1–6. [Google Scholar] [CrossRef]
  41. Orlova, A.; Kemenov, D.; Pet’Kov, V.; Zharinova, M.; Kazantsev, G.; Samoĭlov, S.; Kurazhkovskaya, V. Ultralow and negative thermal expansion in zirconium phosphate ceramics. High Temp. High Press. 2002, 34, 315–322. [Google Scholar] [CrossRef]
  42. Orlova, A.I.; Kemenov, D.V.; Samoilov, S.G.; Kazantsev, G.N.; Pet’Kov, V.I. Thermal expansion of NZP-family alkali-metal (Na, K) zirconium phosphates. Inorg. Mater. 2000, 36, 955–1000. [Google Scholar] [CrossRef]
  43. Anantharamulu, N.; Rao, K.K.; Rambabu, G.; Kumar, B.V.; Radha, V.; Vithal, M. A wide-ranging review on Nasicon type materials. J. Mater. Sci. 2011, 46, 2821–2837. [Google Scholar] [CrossRef]
  44. Suganth, M.; Kumar, N.R.S.; Varadaraju, U.V. Synthesis and leachability studies of NZP and eulytine phases. Waste Manag. 1998, 18, 275–279. [Google Scholar] [CrossRef]
  45. Buvaneswari, G.; Varadaraju, U.V. Low leachability phosphate lattices for fixation of select metal ions. Mater. Res. Bull. 2000, 35, 1313–1323. [Google Scholar] [CrossRef]
  46. Pet’kov, V.; Asabina, E.; Loshkarev, V.; Sukhanov, M. Systematic investigation of the strontium zirconium phosphate ceramic form for nuclear waste immobilization. J. Nucl. Mater. 2016, 471, 122–128. [Google Scholar] [CrossRef]
  47. Sukhanov, M.V.; Pet’kov, V.I.; Firsov, D.V. Sintering mechanism for high-density NZP ceramics. Inorg. Mater. 2011, 47, 674–678. [Google Scholar] [CrossRef]
  48. Tokita, M. Progress of Spark Plasma Sintering (SPS): Method, Systems, Ceramics Applications and Industrialization. Ceramics 2021, 4, 160–198. [Google Scholar] [CrossRef]
  49. Tokita, M. Spark Plasma Sintering (SPS) Method, Systems, and Applications (Chapter 11.2.3). In Handbook of Advanced Ceramics, 2nd ed.; Somiya, S., Ed.; Academic Press: Cambridge, MA, USA, 2013; pp. 1149–1177. [Google Scholar] [CrossRef]
  50. Olevsky, E.A.; Dudina, D.V. Field-Assisted Sintering: Science and Application; Springer: Berlin/Heidelberg, Germany, 2018. [Google Scholar] [CrossRef]
  51. Orlova, A.I.; Koryttseva, A.K.; Kanunov, A.E.; Chuvil’Deev, V.N.; Moskvicheva, A.V.; Sakharov, N.V.; Boldin, M.S.; Nokhrin, A.V. Fabrication of NaZr2(PO4)3-type ceramic materials by spark plasma sintering. Inorg. Mater. 2012, 48, 372–377. [Google Scholar] [CrossRef]
  52. Potanina, E.; Orlova, A.; Nokhrin, A.; Boldin, M.; Sakharov, N.; Belkin, O.; Chuvil’Deev, V.; Tokarev, M.; Shotin, S.; Zelenov, A. Characterization of Nax(Ca/Sr)1-2xNdxWO4 complex tungstates fine-grained ceramics obtained by Spark Plasma Sintering. Ceram. Int. 2018, 44, 4033–4044. [Google Scholar] [CrossRef]
  53. Orlova, A.; Volgutov, V.; Mikhailov, D.; Bykov, D.; Skuratov, V.; Chuvil’Deev, V.; Nokhrin, A.; Boldin, M.; Sakharov, N. Phosphate Ca1/4Sr1/4Zr2(PO4)3 of the NaZr2(PO4)3 structure type: Synthesis of a dense ceramic material and its radiation testing. J. Nucl. Mater. 2014, 446, 232–239. [Google Scholar] [CrossRef]
  54. Papynov, E.; Shichalin, O.; Mayorov, V.; Modin, E.; Portnyagin, A.; Gridasova, E.; Agafonova, I.; Zakirova, A.; Tananaev, I.; Avramenko, V. Sol-gel and SPS combined synthesis of highly porous wollastonite ceramic materials with immobilized Au-NPs. Ceram. Int. 2017, 43, 8509–8516. [Google Scholar] [CrossRef]
  55. 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]
  56. Shichalin, O.O.; Papynov, E.K.; Maiorov, V.Y.; Belov, A.A.; Modin, E.B.; Buravlev, I.Y.; Azarova, Y.A.; Golub, A.V.; Gridasova, E.A.; Sukhorada, A.E.; et al. Spark Plasma Sintering of Aluminosilicate Ceramic Matrices for Immobilization of Cesium Radionuclides. Radiochemistry 2019, 61, 185–191. [Google Scholar] [CrossRef]
  57. Mikhailov, D.A.; Potanina, E.A.; Orlova, A.I.; Nokhrin, A.V.; Boldin, M.S.; Belkin, O.A.; Sakharov, N.V.; Skuratov, V.A.; Kirilkin, N.S.; Chuvil’Deev, V.N. Radiation resistance and hydrolytic stability of Y0.95Cd0.05PO4-based ceramic with the xenotime structure. Inorg. Mater. 2021, 57, 760–765. [Google Scholar] [CrossRef]
  58. Mikhailov, D.; Orlova, A.; Malanina, N.; Nokhrin, A.; Potanina, E.; Chuvil’Deev, V.; Boldin, M.; Sakharov, N.; Belkin, O.; Kalenova, M.; et al. A study of fine-grained ceramics based on complex oxides ZrO2-Ln2O3 (Ln = Sm, Yb) obtained by Spark Plasma Sintering for inert matrix fuel. Ceram. Int. 2018, 44, 18595–18608. [Google Scholar] [CrossRef]
  59. Alekseeva, L.; Nokhrin, A.; Boldin, M.; Lantsev, E.; Murashov, A.; Orlova, A.; Chuvil’Deev, V. Study of the hydrolytic stability of fine-grained ceramics based on Y2.5Nd0.5AlO12 oxide with a garnet structure under hydrothermal conditions. Materials 2021, 14, 2152. [Google Scholar] [CrossRef]
  60. Berendeev, N.N.; Popov, A.A.; Pyaterikova, V.V.; Boldin, M.S.; Nokhrin, A.V.; Chuvil’deev, V.N.; Lantsev, E.A. Modeling of the distribution of thermal fields during spark plasma sintering of alumina ceramics. IOP Conf. Ser. Mater. Sci. Eng. 2019, 558, 012004. [Google Scholar] [CrossRef]
  61. Bernard-Granger, G.; Benameur, N.; Guizard, C.; Nygren, M. Influence of graphite contamination on the optical properties of transparent spinel obtained by spark plasma sintering. Scr. Mater. 2009, 60, 164–167. [Google Scholar] [CrossRef]
  62. Nečina, V.; Pabst, W. Reduction on temperature gradient and carbon contamination in electric current assisted sintering (ECAS/SPS) using a “saw-tooth” heating schedule. Ceram. Int. 2019, 45, 22987–22990. [Google Scholar] [CrossRef]
  63. Hammoud, H.; Garnier, V.; Fantozzi, G.; Lachaud, E.; Tadier, S. Mechanism of carbon contamination in transparent MgAl2O4 and Y3Al5O12 ceramics sintered by Spark Plasma Sintering. Ceramics 2019, 2, 612–619. [Google Scholar] [CrossRef]
  64. Mikhailov, D.A.; Potanina, E.A.; Nokhrin, A.V.; Orlova, A.I.; Yunin, P.A.; Sakharov, N.V.; Boldin, M.S.; Belkin, O.A.; Skuratov, V.A.; Issatov, A.T.; et al. Investigation of microstructure of fine-grained YPO4:Gd ceramics with xenotime structure after Xe irradiation. Ceramics 2022, 5, 237–252. [Google Scholar] [CrossRef]
  65. Nokhrin, A.; Andreev, P.; Boldin, M.; Chuvil’deev, V.; Chegurov, M.; Smetanina, K.; Gryaznov, M.; Shotin, S.; Nazarov, A.; Shcherbak, G.; et al. Investigation of microstructure and corrosion resistance of Ti-Al-V titanium alloys obtained by Spark Plasma Sintering. Metals 2021, 11, 945. [Google Scholar] [CrossRef]
  66. Wang, P.; Yang, M.; Zhang, S.; Tu, R.; Goto, T.; Zhang, L. Suppression of carbon contamination in SPSed CaF2 transparent ceramics by Mo foil. J. Eur. Ceram. Soc. 2017, 37, 4103–4107. [Google Scholar] [CrossRef]
  67. Wang, P.; Huang, Z.; Morita, K.; Li, Q.; Yang, M.; Zhang, S.; Goto, T.; Tu, R. Influence of spark plasma sintering conditions on microstructure, carbon contamination, and transmittance of CaF2 ceramics. J. Eur. Ceram. Soc. 2022, 42, 245–257. [Google Scholar] [CrossRef]
  68. Hagman, L.O.; Kierkegard, P. The crystal structure of NaM2IV (PO4)3; MeIV = Ge, Ti, Zr. Acta Chem. Scand. 1968, 22, 1822–1832. [Google Scholar] [CrossRef]
Figure 1. Regular sintering diagram “temperature-stress-time” for SPS. Na0.7Zr2(PO4)2.7(MoO4)0.3 compound.
Figure 1. Regular sintering diagram “temperature-stress-time” for SPS. Na0.7Zr2(PO4)2.7(MoO4)0.3 compound.
Materials 16 00990 g001
Figure 2. Temperature curves of the shrinkage of Na0.7Zr2(PO4)2.7(MoO4)0.3 powders: Leff—the effective shrinkage, L0—the contribution of the thermal expansion of the setup, L—true shrinkage.
Figure 2. Temperature curves of the shrinkage of Na0.7Zr2(PO4)2.7(MoO4)0.3 powders: Leff—the effective shrinkage, L0—the contribution of the thermal expansion of the setup, L—true shrinkage.
Materials 16 00990 g002
Figure 3. SEM images of the agglomerates (a,c,e) and of the synthesized powders (b,d,f): Ca0.4Zr2(MoO4)0.1(PO4)2.9 (a,b), Na0.5Zr2 (MoO4)0.5(PO4)2 (c,d), Na0.5Zr2 (WO4)0.5(PO4)2 (e,f).
Figure 3. SEM images of the agglomerates (a,c,e) and of the synthesized powders (b,d,f): Ca0.4Zr2(MoO4)0.1(PO4)2.9 (a,b), Na0.5Zr2 (MoO4)0.5(PO4)2 (c,d), Na0.5Zr2 (WO4)0.5(PO4)2 (e,f).
Materials 16 00990 g003aMaterials 16 00990 g003b
Figure 4. TEM images of the powders Ca1-xZr2(PO4)3-x(MoO4)x with x = 0.1 (a,b), x = 0.3 (c,d), x = 0.4 (e,f).
Figure 4. TEM images of the powders Ca1-xZr2(PO4)3-x(MoO4)x with x = 0.1 (a,b), x = 0.3 (c,d), x = 0.4 (e,f).
Materials 16 00990 g004aMaterials 16 00990 g004b
Figure 5. TEM images of the powders Na0.5Zr2(PO4)2.5(XO4)0.5: X = Mo (a,b,e), X = W (c,d,f).
Figure 5. TEM images of the powders Na0.5Zr2(PO4)2.5(XO4)0.5: X = Mo (a,b,e), X = W (c,d,f).
Materials 16 00990 g005
Figure 6. XRD data. Phosphate Na1-xZr2(PO4)3-x(XO4)x compounds, X = Mo (1), W (2), x = 0 (a), 0.1 (b), 0.2 (c), 0.3 (d), 0.4 (e), 0.5 (f).
Figure 6. XRD data. Phosphate Na1-xZr2(PO4)3-x(XO4)x compounds, X = Mo (1), W (2), x = 0 (a), 0.1 (b), 0.2 (c), 0.3 (d), 0.4 (e), 0.5 (f).
Materials 16 00990 g006
Figure 7. XRD data. Phosphate Ca1-xZr2(PO4)3-x(XO4)x compounds, X = Mo (1), W (2), x = 0 (a), 0.1 (b), 0.2 (c), 0.3 (d), 0.4 (e), 0.5 (f).
Figure 7. XRD data. Phosphate Ca1-xZr2(PO4)3-x(XO4)x compounds, X = Mo (1), W (2), x = 0 (a), 0.1 (b), 0.2 (c), 0.3 (d), 0.4 (e), 0.5 (f).
Materials 16 00990 g007
Figure 8. Dependencies of the CSR sizes (DXRD) on the Mo and W contents in the Na1-xZr2(PO4)3-x(XO4)x powders.
Figure 8. Dependencies of the CSR sizes (DXRD) on the Mo and W contents in the Na1-xZr2(PO4)3-x(XO4)x powders.
Materials 16 00990 g008
Figure 9. Dependencies of the unit cell parameters a and c, and of the unit cell volume V on the composition (x) of phosphates Na1-xZr2(PO4)3-x(XO4)x (a,c,e) and Ca0.5(1-x)Zr2(PO4)3-x(XO4)x (b,d,f), X = Mo (a), W (b). The linear fit of the data for the Ca0.5(1-x)Zr2(PO4)3-x(XO4)x compounds was performed without taking into account the data for x = 0.1. The reliability coefficient of linear fit for Ca0.5(1-x)Zr2(PO4)3-x(XO4)x was R2 > 0.9 without taking into account x = 0.1.
Figure 9. Dependencies of the unit cell parameters a and c, and of the unit cell volume V on the composition (x) of phosphates Na1-xZr2(PO4)3-x(XO4)x (a,c,e) and Ca0.5(1-x)Zr2(PO4)3-x(XO4)x (b,d,f), X = Mo (a), W (b). The linear fit of the data for the Ca0.5(1-x)Zr2(PO4)3-x(XO4)x compounds was performed without taking into account the data for x = 0.1. The reliability coefficient of linear fit for Ca0.5(1-x)Zr2(PO4)3-x(XO4)x was R2 > 0.9 without taking into account x = 0.1.
Materials 16 00990 g009
Figure 10. Temperature dependencies of the unit cell parameters a and c, and of the unit cell volume V of phosphates of Na1-xZr2(PO4)3-x(XO4)x, X = Mo (a,c,e), W (b,d,f); x = 0, 0.1, 0.2, 0.3, 0.4, 0.5.
Figure 10. Temperature dependencies of the unit cell parameters a and c, and of the unit cell volume V of phosphates of Na1-xZr2(PO4)3-x(XO4)x, X = Mo (a,c,e), W (b,d,f); x = 0, 0.1, 0.2, 0.3, 0.4, 0.5.
Materials 16 00990 g010aMaterials 16 00990 g010b
Figure 11. Temperature dependencies of the unit cell parameters a and c, and of the unit cell volume V of phosphates of Ca1-xZr2(PO4)3-x(XO4)x, X = Mo (a,c,e), W (b,d,f); x = 0, 0.1, 0.2, 0.3.
Figure 11. Temperature dependencies of the unit cell parameters a and c, and of the unit cell volume V of phosphates of Ca1-xZr2(PO4)3-x(XO4)x, X = Mo (a,c,e), W (b,d,f); x = 0, 0.1, 0.2, 0.3.
Materials 16 00990 g011aMaterials 16 00990 g011b
Figure 12. Dependencies of the thermal expansion parameters (αa, αc, αav, β, and Δα) on the compositions (x) of phosphates Na1-xZr2(PO4)3-x(XO4)x, X = Mo (a), W (b).
Figure 12. Dependencies of the thermal expansion parameters (αa, αc, αav, β, and Δα) on the compositions (x) of phosphates Na1-xZr2(PO4)3-x(XO4)x, X = Mo (a), W (b).
Materials 16 00990 g012
Figure 13. Dependencies of the thermal expansion parameters (αa, αc, αav, β, and Δα) on the compositions (x) of phosphates Ca0.5(1-x)Zr2(PO4)3-x(XO4)x, X = Mo (a), W (b).
Figure 13. Dependencies of the thermal expansion parameters (αa, αc, αav, β, and Δα) on the compositions (x) of phosphates Ca0.5(1-x)Zr2(PO4)3-x(XO4)x, X = Mo (a), W (b).
Materials 16 00990 g013
Figure 14. SPS diagrams L(T) (a,b) and S(T) (c,d) for the Na1-xZr2(PO4)3-x(XO4)x ceramics, X = Mo (a,c), W (b,d).
Figure 14. SPS diagrams L(T) (a,b) and S(T) (c,d) for the Na1-xZr2(PO4)3-x(XO4)x ceramics, X = Mo (a,c), W (b,d).
Materials 16 00990 g014
Figure 15. XRD curves of the ceramic samples of phosphate–tungstate Na1-xZr2(PO4)3-x(XO4)x ceramics, X = W (a), Mo (b), with x = 0.5: W1–W3—the labels of three phosphate–tungstate satellite ceramic specimens, M1–M7—the labels of seven phosphate-molybdate ones.
Figure 15. XRD curves of the ceramic samples of phosphate–tungstate Na1-xZr2(PO4)3-x(XO4)x ceramics, X = W (a), Mo (b), with x = 0.5: W1–W3—the labels of three phosphate–tungstate satellite ceramic specimens, M1–M7—the labels of seven phosphate-molybdate ones.
Materials 16 00990 g015
Figure 16. Microstructure of Na1-xZr2(PO4)3-x(MoO4)x ceramics with x = 0.1 (a,b) and x = 0.5 (c,d). SEM images of the fractures of sintered ceramic samples (b,d).
Figure 16. Microstructure of Na1-xZr2(PO4)3-x(MoO4)x ceramics with x = 0.1 (a,b) and x = 0.5 (c,d). SEM images of the fractures of sintered ceramic samples (b,d).
Materials 16 00990 g016
Figure 17. Microstructure of Na1-xZr2(PO4)3-x(WO4)x ceramics with x = 0.1 (a,b) and x = 0.5 (c,d). SEM images of the fractures of sintered ceramic samples (b,d).
Figure 17. Microstructure of Na1-xZr2(PO4)3-x(WO4)x ceramics with x = 0.1 (a,b) and x = 0.5 (c,d). SEM images of the fractures of sintered ceramic samples (b,d).
Materials 16 00990 g017
Table 1. Crystallographic characteristics of the phosphates Na1-xZr2(PO4)3-x(XO4)x (denoted as “NZP”) and Ca0.5(1-x)Zr2(PO4)3-x(XO4)x (denoted as “CZP”).
Table 1. Crystallographic characteristics of the phosphates Na1-xZr2(PO4)3-x(XO4)x (denoted as “NZP”) and Ca0.5(1-x)Zr2(PO4)3-x(XO4)x (denoted as “CZP”).
Xxa, Åc, ÅV, Å3
NZPCZPNZPCZPNZPCZP
-08.799(4)8.784(6)22.826(7)22.736(0)1530.6(7)1519.4(7)
Mo0.18.811(8)8.792(0)22.856(6)22.762(2)1536.9(9)1523.7(7)
0.28.825(5)8.801(2)22.882(7)22.777(7)1543.5(3)1528.0(2)
0.38.833(4)-22.904(9)-1547.8(0)-
0.48.851(9)-22.921(5)-1555.4(2)-
0.58.891(7)-22.897(1)-1567.7(8)-
W0.18.821(6)8.813(5)22.862(2)23.079(8)1540.7(7)1552.6(0)
0.28.827(5)8.815(2)22.935(1)22.949(0)1547.7(9)1544.3(9)
0.38.833(3)8.827(4)22.989(5)22.996(8)1553.4(7)1551.9(0)
0.48.857(1)-23.046(1)-1565.7(0)-
0.58.881(3)-23.053(7)-1574.8(0)-
Table 2. Thermal expansion coefficients [in °C−1] of phosphates Na1-xZr2(PO4)3-x(XO4)x (denoted as “NZP”) and Ca0.5(1-x)Zr2(PO4)3-x(XO4)x (denoted as “CZP”).
Table 2. Thermal expansion coefficients [in °C−1] of phosphates Na1-xZr2(PO4)3-x(XO4)x (denoted as “NZP”) and Ca0.5(1-x)Zr2(PO4)3-x(XO4)x (denoted as “CZP”).
Xxαa·106αc·106αav·106β·106Δα·106
NZPCZPNZPCZPNZPCZPNZPCZPNZPCZP
-0−4.20−2.9620.3710.603.991.5611.824.6024.5613.56
Mo0.1−4.43−2.7319.8210.103.661.5510.904.7124.2512.83
0.2−3.63−2.6120.509.574.411.4513.144.3024.1212.18
0.3−3.62-20.20-4.32-12.75-23.82-
0.4−4.29-17.49-2.97-8.75-21.79-
0.5−2.81-15.64-3.34-9.92-18.45-
W0.1−5.10−0.5720.036.243.281.709.755.0025.136.81
0.2−3.51−2.5017.0012.333.332.449.997.2620.5214.83
0.3−2.94−2.2615.0910.803.072.099.236.2918.0413.06
0.4−2.71-11.85-2.14-6.45-14.56-
0.5−3.71-11.65-1.41-4.23-15.36-
Table 3. Characteristics of ceramic samples of phosphates Na1-xZr2(PO4)3-x(XO4)x.
Table 3. Characteristics of ceramic samples of phosphates Na1-xZr2(PO4)3-x(XO4)x.
XxVh, oC/minTs, °Ctsps, minρ, %Hv, GPaKIC, MPa·m1/2
Mo0.15092012.5897.553.6 ± 0.51.0 ± 0.3
0.293213.0898.424.5 ± 0.81.1 ± 0.5
0.3111016.50100.075.3 ± 0.50.9 ± 0.3
0.497513.58100.065.1 ± 0.81.1 ± 0.6
0.586111.50101.665.3 ± 0.51.4 ± 0.3
W0.150119018.0897.705.3 ± 0.51.3 ± 0.3
0.2110016.8398.755.6 ± 0.61.6 ± 0.3
0.3111017.5898.595.6 ± 0.51.5 ± 0.3
0.4106515.67100.205.5 ± 0.41.1 ± 0.3
0.5106515.08100.705.6 ± 0.61.1 ± 0.6
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Karaeva, M.E.; Savinykh, D.O.; Orlova, A.I.; Khainakov, S.A.; Nokhrin, A.V.; Boldin, M.S.; Garcia-Granda, S.; Murashov, A.A.; Chuvil’deev, V.N.; Yunin, P.A.; et al. (Na, Zr) and (Ca, Zr) Phosphate-Molybdates and Phosphate-Tungstates: I–Synthesis, Sintering and Characterization. Materials 2023, 16, 990. https://doi.org/10.3390/ma16030990

AMA Style

Karaeva ME, Savinykh DO, Orlova AI, Khainakov SA, Nokhrin AV, Boldin MS, Garcia-Granda S, Murashov AA, Chuvil’deev VN, Yunin PA, et al. (Na, Zr) and (Ca, Zr) Phosphate-Molybdates and Phosphate-Tungstates: I–Synthesis, Sintering and Characterization. Materials. 2023; 16(3):990. https://doi.org/10.3390/ma16030990

Chicago/Turabian Style

Karaeva, M. E., D. O. Savinykh, A. I. Orlova, S. A. Khainakov, A. V. Nokhrin, M. S. Boldin, S. Garcia-Granda, A. A. Murashov, V. N. Chuvil’deev, P. A. Yunin, and et al. 2023. "(Na, Zr) and (Ca, Zr) Phosphate-Molybdates and Phosphate-Tungstates: I–Synthesis, Sintering and Characterization" Materials 16, no. 3: 990. https://doi.org/10.3390/ma16030990

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