Next Article in Journal
3D Printing Ceramics—Materials for Direct Extrusion Process
Next Article in Special Issue
NMR Spectra Particularities in LiNbO3 Crystals with a Near-Stoichiometric Composition
Previous Article in Journal
Combining Pultrusion with Carbonization: Process Analysis and Material Properties of CFRP and C/C
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multifunctional Materials Based on the Solid Solutions of the (Na, K, Cd0.5)NbO3 (KNN-Cd) System Modified with Rare Earth Elements: Crystal Structure, Microstructure, and Macroresponses

1
Research Institute of Physics, Southern Federal University, Rostov-on-Don 344002, Russia
2
Tananaev Institute of Chemistry—Subdivision of the Federal Research Centre, Kola Science Centre of the Russian Academy of Sciences, Apatity 184209, Russia
3
Institute of High Technology and Piezo Technic, Southern Federal University, Rostov-on-Don 344002, Russia
4
Faculty of Physics, Southern Federal University, Rostov-on-Don 344002, Russia
*
Author to whom correspondence should be addressed.
Ceramics 2023, 6(1), 342-363; https://doi.org/10.3390/ceramics6010021
Submission received: 27 December 2022 / Revised: 27 January 2023 / Accepted: 29 January 2023 / Published: 1 February 2023

Abstract

:
Ceramic samples of polycomponent solid solutions (SSs) Na0.90K0.05Cd0.05NbO3 (Cd-modified KNN) and [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3, where REE = La, Pr, Tb, Dy, Ho, were obtained by the uniaxial hot pressing (UHP) method. The crystal lattice structures, morphological features of the microstructure, and dielectric and thermophysical properties of these ceramics have been investigated. For the first time, their strength characteristics (Young’s modulus) and the critical stress intensity factor of the mode I Kic have been estimated. Photoluminescent properties have been compared in SSs [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3 in the visible wavelength range.

1. Introduction

Almost every decade, starting from the 1950s, marks a new vector of development in functional materials science. Ferroelectricity was discovery in BaTiO3 in the 1950s, and the materials based on it appeared [1,2,3,4]. Compositions based on the binary PZT-system of (Pb(Zr,Ti)O3) were studied in the 1960s [5,6], including the modified ones [7,8,9,10]. Ternary systems of the solid solutions (SS) [11,12] were the most popular materials studied in the 1970s. In the 1980s, systems became the four- and five-component SS [13,14,15,16]. A study and application of lead-free materials based on ((Na,Li)NbO3, (Na, K)NbO3 (KNN) started in the 1990s due to ecological matters. The Restriction of Hazardous Substances Directive (RoHR) was officially adopted in 2003 in the European Union; it forbade use of lead in electronic devices [17], thus lead-free materials became very relevant. The multicomponent lead-free KNN niobate materials prevailed in the 2000s [18,19,20,21]. Saito et al. have proved in 2004 [22] that a KNN-based textured ceramic has a very high d333 value (d333 = 416 pC/N). The authors of [23] have revealed a twice larger d333 value in this material. Lead-free KNN materials were intensely modified and became multi-element complex compounds in the 2010s [24,25]. The 2020s are marked by the modification of the KNN with rare earth elements (REEs) and the development of the new lead-free complex compositions, including the compounds with fundamentally different properties [26,27,28,29].
Thus, current material science develops in several directions: technology becomes greener; basic compositions of functional materials are complicated; non-isostructural components are incorporated into the initial systems; matrices are doped with exotic modifiers. This is made to establish and integrate in materials various macro-responses, such as ferroelectric, piezoelectric, thermophysical, mechanical, luminescent, etc. All the above-mentioned modifications expand application possibilities of such multifunctional materials.
This work researches solid solutions (SSs) with the composition Na0.90K0.05Cd0.05NbO3. The SSs belong to the class of complex ferroelectric perovskites [30] of the ABO3 type. The Na0.90K0.05Cd0.05NbO3 SS structure contains cuboctahedron voids (AO12 polyhedra). Voids can be occupied by several types of cations [31]. At this, sizes of cations can be other than the sizes of the polyhedra they occupy. As a result, interatomic bonds strain, and perfect perovskite subcells are significantly distorted. This leads to deformation and rotation of anionic octahedra in NaNbO3-based SSs [32,33]. A wide variety of both crystal lattice distortions and physical properties depend on the temperature and composition in NaNbO3-based SSs, which is rare among perovskite-type materials. The initial compound NaNbO3 exhibits a number of structural phase transitions [34,35]. This crystalochemical feature causes the above-mentioned dependence. NaNbO3-based SS is characterized by the concentration and temperature phase transitions of the displacement type. Some of the phase transitions are associated with a change in the type of dipole ordering or with morphotropic regions of coexistence of phases characterized by differently ordered structure distortions [36,37]. The characteristics of morphotropic regions, the structures of individual phases, and their properties (including mechanical ones) depend on the microhomogeneity and the degree of structural ordering of these SSs. This is the reason why SSs are sensitive to the conditions of sample preparation.
Mechanical (for example, elastic) properties are crucial for the study of solids. Young’s modulus characterizes interatomic interactions and macroscopic anisotropy of a solid. Elastic constants reveal many fundamental features of the interaction of particles in matter; the constants determine the P and T regions of phonon instability of the crystal lattice and much more.
The aim of this work is to expand the number of known properties of ceramics based on NaNbO3. In order to achieve this aim, we have synthesized multicomponent ceramic ferroelectric Na0.90K0.05Cd0.05NbO3 SSs doped with 5 mol% REE (La, Pr, Tb, Dy, Ho). We have studied their structure, microstructure, dielectric, thermophysical, and optical (luminescent) characteristics. The performed study has widened potential application of [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3 SSs compared to the application range of the initial undoped Na0.90K0.05Cd0.05NbO3 SSs.

2. Materials and Methods

2.1. Materials

The objects of study were the SSs of the composition Na0.90K0.05Cd0.05NbO3 and [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3, where REE = La, Pr, Tb, Dy, Ho. SSs were synthesized in accordance with the specified composition. The following initial components were mixed: sodium bicarbonate (NaHCO3, 99.9 (not more than 10−5 of 27 impurities), Komponent-reaktiv Ltd., Russia), potassium bicarbonate (KHCO3, 99.9%, Solikamsk Magnesium Plant, Russia), niobium (Nb2O5, 99.9%, Solikamsk Magnesium Plant, Russia), and cadmium oxides (CdO, 99%). The lanthanum (La2O3, 99.9, Himkraft, Russia), praseodymium (Pr2O3, 99.9, Himkraft, Russia), terbium (Tb2O3, 99.9, Himkraft, Russia), dysprosium (Dy2O3, 99.9, Himkraft, Russia), and holmium (Ho2O3, 99.9, Himkraft, Russia) oxides were used as the modifying dopants.
The samples were obtained by the two-stage solid-phase synthesis (Tsynt.1 = 1020 K, τ = 5 h; Tsynt.2 = 1070 K, τ = 5 h). After that the Ø10 mm and 8–10 mm high blanks were sintered and treated by uniaxial hot pressing (UHP). Pressure was applied externally using the modernized installation for uniaxial hot pressing [38]. The optimal sintering modes for the UHP technology were determined: Tsint = 1460 K, P = 230 kg/cm2, and τsint = 60 min. The pressure was applied at room temperature and maintained at the stages of the temperature rise and isothermal holding. The applied pressure was removed during cooling. The preparation modes were optimized for each composition: the sintering temperature varied in a series of samples. Phase composition of the synthesized and sintered blanks was controlled with the XRD. Additionally, the microstructure was analyzed and the ceramics density was determined. The selected technology regulations were obtained on the basis of these results. The technology ensured the absolute purity or minimization of the ballast phases in synthesized powders and sintered ceramics. Ballast phases cannot be completely excluded within the applied method of the SSs preparation. The technology also ensured the high density, a sufficiently high quality of their microstructure, and the integrity and intactness of the ceramic frame.
The sintered ceramic blanks were mechanically treated: cut along the plane, grinded and polished along the flat surfaces and ends. As a result, samples were Ø10 mm × 1 mm. Each composition provided 8–10 sample pieces. Before the metal spraying, the samples were calcined at Tcalc. = 770 K for 0.5 h to remove the organic residues and degrease the surfaces in order to increase the adhesion of the metal coating to the ceramics. The electrodes were applied by the double firing of a silver-containing paste onto the flat sample surfaces at 1070 K for 0.5 h.

2.2. Research

The XRD studies were performed by the powder diffraction on a DRON-3 diffractometer (NPP Burevestnik, St. Petersburg, Russia, 2012) with an IR-2 X-ray interferometer (Radicon Ltd., St. Petersburg, Russia) using Co radiation. The IR-2 was connected to a computer equipped with special programs. Parameters of the perovskite cell and the uniform deformation parameter δ were calculated. The deformation parameter characterizes the spontaneous deformation value. Calculation formulas corresponded to the monoclinic symmetry of the crystal lattice. Formulas were simplified, taking into account the small deviation of the perovskite cell from the cubic one [38]:
d h k l = a sin β N ( 1 + k 2 N y + l h N cos β )
δ = a 1 cos β a ¯ 1
where N = h2 + k2 + l2; h, k, l—the diffraction indices; y = (b/a sin β)—1; a, b, β—the parameters of the perovskite cell; a ¯ V 3 —the average parameter. The monoclinic (M) unit cell parameters were calculated from the diffraction peaks 200, 020, and 22 2 ¯ . The measure errors for the structural parameters were as follows: ∆a = ∆b = ∆c = ±0.004 Å, Δβ = ±0.05°, and ∆V = ± 0.10 Å3.
The experimental density of the samples was measured by hydrostatic weighing in octane. The XRD density was calculated by the formula: ρx-ray = 1.66·M/V, where M is the molecular weight per cell, and V is the volume of the cell. Integral density was measured for every sample obtained from the same sintered blank. This confirmed a high homogeneity of the samples under study and the reliability of the results obtained in the work.
The microstructure of the ceramic chips was studied by a JSM-6390L scanning electron microscope (JEOL Ltd., Tokyo, Japan, 2003) with an AZtecLive microanalyzer system (Oxford Instruments, High Wycombe, Great Britain). The microscope resolution was up to 1.2 nm at an accelerating voltage of 30 kV for the image in the secondary electrons, the accelerating voltage range was from 0.5 to 30 kV, the magnification was from ×10 to ×1,000,000, and the beam current was up to 200 nA. The study of the morphology was performed using a VK-9700 color laser scanning 3D microscope (KEYENCE, Itasca, IL, USA) and a TM1000 electron microscope (Hitachi, Tokyo, Japan).
The temperature dependences of real and imaginary parts of the relative complex permittivity were obtained using a measuring stand based on an Agilent 4980 A LCR-meter (Agilent Technologies, USA, Santa Clara, 2008). The temperature was measured at T = 300–900 K and in the frequency range f = 25–1 × 106 Hz. By real and imaginary parts of the relative complex permittivity we mean ε*/ε0 = ε’/ε0 − iε’’/ε0, where ε0 = 8.75·10−12 F/m—the dielectric constant of the samples.
Low-temperature dielectric measurements were carried out in the temperature range T = 10–300 K at frequencies f = 20–1 × 106 Hz using a Wayne Kerr 6500 B precision impedance analyzer (Wayne Kerr Electronics, UK, Bognor Regis, 2006). The samples were cooled using a CCS-150 Cryogenics closed-type helium refrigeration cryostat (Lake Shore Cryotronics Inc., Westerville, OH, USA). The temperature was controlled using a Model 331 temperature controller (Lake Shore Cryotronics Inc., Westerville, OH, USA). We were able to maintain the set temperature with an accuracy of ±0.01 K. During measurements, the samples were placed in the vacuum chamber of the cryostat; the vacuum was created by a BOC turbomolecular pump (Edwards, Stockholm, Sweden).
The thermal expansion (ΔL/L) and the linear coefficient of the thermal expansion (α) were studied using a specially designed measuring installation based on a MICRON-02 device for checking the end rods (“TOT” Enterprise, Russia, Yaroslavl, 2010), and an nanovoltmeter/microohmmeter Agilent 34420A (Agilent Technologies, USA, Santa Clara, 2012). The sensitivity to the displacement was 10−8 m, and the measurement error was 4%. The temperature in the chamber was changed at a rate of ΔT = 1 K/min.
The mechanical properties of the ceramics were investigated by the contact method on a NanoSkan probe microscope–nanohardness tester (FSBI TISNCM, Troitsk, Russia).
The photoluminescence (PL) spectra of the studied ceramic samples were recorded using a SOL SL-100M spectrograph (SOL instruments, Belarus, Minsk) with an FLI ML 1107 Black Illuminated CCD detector (Hamamatsu PHOTONICS K.K., Japan, Hamamatsu City) in the spectral range 380–800 nm. The entrance slit of the monochromator was 0.4 mm. A He-Cd continuous laser (Kimmon KOHA, Japan, Fukushima) set to λ0 = 325 nm and P = 15 mW was used as a luminescence excitation source. Each PL spectrum was corrected to the background signal.

3. Results and Discussion

3.1. X-Ray Study

The initial Na0.90K0.05Cd0.05NbO3 SS crystallizes in a perovskite-type structure; it has the rhombic symmetry with a monoclinically distorted perovskite cell. The correlations between the parameters of the monoclinic cell a = c > b and α = γ < β, and the rhombic cell A, B, and C are as follows: A = 2a cos β/2, B = 2b, and C = 2a sin β/2. Figure 1 shows the XRD patterns of the initial Na0.90K0.05Cd0.05NbO3 (curve 1) and the modified [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3, REE = La, Pr, Tb, Dy, Ho (2, 3, 4, 5, 6, respectively) SSs. The peaks are characterized by the data available from [39].
The XRD pattern of the Na0.90K0.05Cd0.05NbO3 SS shows the splitting of the multiplets corresponding to the monoclinic distortion of the perovskite cell. The basic NaNbO3 compound in Figure 1a (1) corresponds to card (JCPDS 14-603). This card was used because KNbO3 and CdNb2O6 are only 5% in the base SS. A very weak superstructural peak (marked with an asterisk) indicates a doubling of the b axis.
The XRD patterns of the modified SSs show that the modification with the La3+ and Pr3+ ions changes the crystal lattice symmetry. The multiplet 200, 020 becomes almost a single peak, and the symmetry increases to the pseudocubic (PsC) one (the XRD patterns 2, 3, Figure 1). The absence of the impurities, the change in the symmetry, and the slight increase in the volume of the perovskite cell (Table 1) indicate that the large-sized La3+ and Pr3+ ions successfully incorporate into the crystal lattice of the initial SS. It should be noted that an increase in the cell volume is possible only with the substitution (La3+, Pr3+)→Na+ (1.04, 1.00, and 0.98, respectively): the ionic radius of K+ (1.33) is much larger than the radii of La3+ and Pr3+; cadmium and these REEs have a large difference in the electronegativities (Cd = 1.69, La = 1.1, and Pr = 1.13), which prevents them from replacing cadmium [40]. The REEs Tb, Dy, and Ho do not completely incorporate into the crystal lattice of the initial SS, but form the impurity LnNbO4 compounds of the monoclinic fergusonite type. As the ionic radii of the REEs decrease in the row Tb→Dy→Ho, the intensity of the XRD peaks of the impurity compound in the XRD patterns 4–6 (Figure 1) increases, and the symmetry of the SS again becomes monoclinic. A single XRD peak 200 on the XRD patterns 2 and 3 (Figure 1) is again transformed into the multiplet 200, 020 (Figure 1: patterns 4, 5, and 6). The volume of the perovskite unit cell (Table 1) decreases in accordance with a decrease in the ionic radii of REEs.
Thus, the precipitation of medium-sized REE ions Tb, Dy, and Ho into the impurity, the return of monoclinic distortion, and a slight decrease in the volume of the perovskite cell indicate that these REEs do not completely incorporate in the original SS. In this case, the profile of the diffraction peak (200) in XRD pattern 5 (Figure 1) and the smallest cell volume (Table 1) indicate that among medium-sized REEs, Dy incorporates into the crystal lattice of the initial SS easier and forms the most homogeneous SS.
Table 1 shows the symmetry and parameters of the perovskite cell of the initial and modified with the REEs SSs based on Na0.9K0.05Cd0.05NbO3, as well as the experimental and relative density of the ceramic samples.
The relative density above 100% can be due to the fact that the measurement of the experimental density in the samples containing the second phase cannot separate the contribution of each phase to the density of the bulk sample. Qualitatively, this can be explained by the fact that the heavy REE ions do not incorporate into the SS; they form instead the LnNbO4 impurity compound. The compound has a much higher density (7.0–7.3 g/cm3) than the initial SS. Therefore, the average experimental density of the sample is overestimated. Since the X-ray density is calculated for a stoichiometric SS, the relative density of the sample can exceed 100%.

3.2. Microstructure

The main characteristic of the microstructure of the ceramics is the average grain size D ¯ . Its value is in a good agreement with the XRD data. Thus, an increase in δ in the samples with Tb, Dy, and Ho leads to a sharp refinement of the morphology due to the existing inverse dependence of D ¯ on δ ( D ¯ ~1/δ [41]). The fact that the compositions with a large δ have a smaller crystallite size can be explained by an increase in the internal stresses in the ceramics with an increase in the spontaneous deformation. It is these stresses that inhibit the growth of the crystallites [41].
An additional factor, which, in our opinion, is also associated with the formation of the fine-grained structures in these samples, as well as in La- and Pr-containing ones, may be the polymorphism of the initial reagents—the REE (Ln2O3) oxides. The oxides undergo a series of the phase transformations during the heat treatment of the charge. At low temperatures (~400 K), oxides interact with water and form Ln(OH)3 hydrates; at higher temperatures (~470–1170 K) oxides decompose [42]. These circumstances are associated with the Hedvall effect—an increase in the reactivity of the solids during or as a result of the polymorphic transformations [43]. This, in turn, causes the simultaneous formation of many centers of the primary recrystallization with the reduced mass capacity; as a result, smaller crystallites grow from each nucleus, forming a new grain pattern. It should be noted that impurities can act as such crystallization centers; the impurities form during the modification of the SS with the low-sized REEs (Tb, Dy, Ho).
As the ionic radii of the REE decreases in the row La→Ho, surface charge density increases; thus, hydration energy on cations increases. Hydration energy is the energy released when interacting with water molecules. As a result, the most significant refinement of the grain structures should be observed in the samples with medium-sized Tb, Dy, and Ho; and they were observed.
In the initial and modified samples, grains crystallize in the form of almost regular geometric figures: parallelepipeds or plates for larger grains; almost regular cubes for smaller grains (Figure 2). This indicates that the crystallization takes place in in the presence of the liquid phases, not in the solid phase [41]. The unreacted initial reagents with low melting points and low-melting eutectics in the charges of alkali metal niobates can serve as sources of liquid phases: Na2O with Tmelt = 1405 K; K2O with Tmelt = 1013 K; CdO with Tmelt = 1173 K; Na2CO3 with Tmelt = 1127 K; K2CO with Tmelt = 1164 K; Li2CO3 with Tmelt = 1005 K; NaOH with Tmelt = 596 K; KOH with Tmelt = 678 K; NaNbO3 with Tmelt = 1695 K; KNbO3 with Tmelt = 1373 K, etc.
The ceramic sample of the initial composition Na0.9K0.05Cd0.05NbO3 is quite different from REE-modified ones (Figure 2a). Its microstructure has no separate grains. The surface of the initial ceramics has a stepped-layered structure fragment (circled by the dashed line on the Figure 2a), which is somewhat unusual. This is probably a manifestation of the linear stacking faults that contain dislocation motives with the steps. The steps are the result of the precipitation of the excess vacancies. The vacancies appear due to the high volatility of the alkali metals; the other reason for vacancies’ appearance is two parts of the stacking faults moving in different directions. A set of such periodically arranged dislocations forms both grain boundaries and ensembles of grain boundaries in the considered SSs. Usually, dislocation motives with the steps are destroyed during recrystallization [42]; in our case they are retained during annealing after the plastic deformation. This, apparently, is a consequence of the externally applied pressure during the sintering of SSs, which fixes these stacking faults.
A similar situation, but on a smaller scale, takes place in the samples with Tb, Dy, and Ho (Figure 2c–e). The effect is absent from the La-modified SSs (Figure 2b). Differential curves of ceramic grain size distribution shown on insets of Figure 2b–f demonstrate that microstructure grains have sizes 0.15–2.5 μm. The average grain size (from 400 to 550 nm) increases in the REE-doped SSs in the dopant row La → Ho. Ceramic grains often do not have clear boundaries and the adhesion between crystallites is quite high. Pores are almost absent.
The observed facts confirm the earlier statement: the large-sized (La and Pr) REEs incorporate into the base crystal lattice completely and stabilize the lattice; the medium-sized Tb, Dy, and Ho on the contrary incorporate into the initial structure only partially. As a result, the lattice does not even change its symmetry.

3.3. Dielectric Properties

Figure 3 and Figure 4 shows the dependences of ε′/ε0 and ε″/ε0 on the temperature at the various frequencies of the alternating electric field obtained in the studied samples cooling mode.
The ceramics of the initial SS composition Na0.90K0.05Cd0.05NbO3 has two peaks of the permittivity constant (Figure 3a). They correspond to the transition from the orthorhombic (O) phase to the tetragonal (T) one (TO-T), and from the tetragonal phase to the cubic one (TT-C) [44]. The dispersion of the dielectric constant begins to manifest itself only above TT-C (Figure 3a).
The introduction of REE dopants with R ≥ 1.0 Å (Pr and La in Figure 3b,c) shifts TO-T and TT-C to the low-temperature region [45]; the same happens with the convergence of TO-T and TT-C upon doping with REE R < 0.92 Å (Tb, Dy, and Ho in Figure 4), which may be due to a decrease in the lattice stability [46]. The shift might be caused by the increase in the disorder in the case of heterovalent substitution in the A-sublattice. In our case, when REEs with R ≥ 1.0 Å are introduced, one can see the following: TC~410 K for La (Figure 3b) and TC~340 K for Pr (Figure 3c).
At R < 0.92 Å (Tb, Dy, and Ho), TC is in the range (500–550) K (Figure 4). The temperature range is so wide in SSs modified with Dy and Ho due to the formation of two small maxima of ε′/ε0 forming together a quasi-plateau region (Figure 4b,c). The latter may be due to the manifestation of the O→T and T→C phase transitions in the base compounds, which are shifted towards each other and form the indicated anomaly [44]. However, this phenomenon requires additional research. In the paraelectric region, ε′/ε0 and ε″/ε0 significantly increase with increasing the temperature (Figure 4). This might be caused by the increasing contribution of the through electrical conductivity to ε′/ε0 and ε″/ε0 dependences. The through electrical conductivity is most likely caused by the redox processes in the ceramics, provoked by the presence of the Nb ions in their compositions; Nb has a variable valence.
The introduction of the REEs blurs ε′/ε0 and ε″/ε0 maxima at the TT-C transition in a quite wide temperature range (Figure 3 and Figure 4, insets in column 1). Thereby, only the unmodified SS with the Na0.9K0.05Cd0.05NbO3 composition (the inset in Figure 3(a1)) in the paraelectric phase obeys the empirical Curie-Weiss law, ε′/ε0 = C/(T−Tc), where the Curie-Weiss constant C = 1.27·105 K, and the Curie-Weiss temperature Tc = 633 K. The measuring field frequency of 520 kHz was chosen to eliminate the contribution of the space charge polarization at high temperatures.
To confirm the possible shift of TO-T and TT-C to different temperature ranges depending on the REE size, we additionally studied ε′/ε0 and ε″/ε0 in the range (10–325) K (Figure 5 and Figure 6). It has been established that the formation of a quasi-maximum at ~245 K is typical for the basic SS Na0.9K0.05Cd0.05NbO3 (Figure 5a). The quasi-maximum may be explained by the TR→TO transition in the basic system (Na, K)NbO3 [47]. A similar effect was observed in the (K0.48Na0.52)(Nb1-xSbx)O3 system [48]. A clearly pronounced hysteresis of the permittivity ε′/ε0 manifests at T~(200–320) K (inset on Figure 5(a1)), while ε″/ε0 decreases without any anomalies in the entire temperature range. However, a weak dispersion ε′/ε0 and ε″/ε0 is observed at T~(220–320) K, while it is almost absent above 320 K. The introduction of large-sized La (Figure 5(b1)) suppressed the indicated ε′/ε0 anomaly, i.e., any inflection points are absent. However, it is necessary to note a weak hysteresis at T~(200–320) K (inset in Figure 5(b1)). The dispersion of ε′/ε0 and ε″/ε0 remains unchanged at T < 200 K of La-modified Na0.9K0.05Cd0.05NbO3 (Figure 5b). At T > 200 K a rapid increase in (ε′/ε0, ε″/ε0)(T) is observed: the lower is the frequency of the measuring electric field, the stronger the increase is. SS containing Pr (Figure 5c) is characterized by an increase in ε′/ε0 and a decrease in ε″/ε0 at T~(10–260). An inflection point forms at T~260 K. After that ε′/ε0(T) of Pr-modified Na0.9K0.05Cd0.05NbO3 demonstrates a plateau-like area at T~(260–320) K and ε″/ε0(T) grows (Figure 5c). A weak hysteresis is present in the dependences (ε′/ε0 and ε″/ε0)(T) up to the inflection point T = 260 K and increases at T > 260 K (Figure 5c). The dispersion of these dependences is observed over the entire temperature range of Pr-modified Na0.9K0.05Cd0.05NbO3 (Figure 5c).
Tb-modified SS has ε′/ε0 and ε″/ε0 maxima in the range (150–320) K of a relaxation nature: ε′/ε0 and ε″/ε0 shift to the region of higher temperatures with an increase in the frequency of the measuring field (Figure 6a). The introduction of small Dy and Ho ions (Figure 6b,c) caused a shift of the maximum observed in the SS Na0.9K0.05Cd0.05NbO3 (Figure 5a) to the low-temperature region (T~200 K). In Dy- and Ho-modified Na0.9K0.05Cd0.05NbO3 the hysteresis decreases to this region, and dispersion of (ε′/ε0, ε″/ε0)(T) stays almost unchanged over the entire low-temperature range under study (Figure 6b,c).
Considering the significant blurring of the TT-C transition, the modified Curie-Weiss law was used to characterize it [49]:
1 ε 1 ε m = ( T T m ) γ C ,   ( T > T m ) ,  
where εm—the dielectric constant at Tm, C is assumed to be the constant, and γ—the degree of the blurring, which can be in the range from 1 for a normal ferroelectric to 2 for an ideal relaxor ferroelectric [50,51]. Figure 7 shows the dependences of log(1/ε′−1/εm) on log(TTm) obtained at a frequency of 520 kHz. All the samples, except those containing Ho and Dy, are characterized by a linear dependence. The slope of the approximating straight line was used to determine the parameter γ [50]. It was found that γ~1.7 upon the modification of the initial SS with the REEs = La, Pr, and Tb. The obtained data indicate a diffuse phase transition. The shift of the maximum of the dielectric constant downward in the temperature indicates an increase in the conductivity in the studied SSs.

3.4. Thermophysical Properties

Figure 8, Figure 9 and Figure 10 show the dependences of ε′/ε0 (the cooling mode), α, and ∆L/L on the temperature of the investigated ceramic SSs.
It was established that in the initial Na0.90K0.05Cd0.05NbO3 SS, ∆L/L(T), both upon heating and cooling, the slope changes for this dependence in the neighborhoods of TO-T and TT-C (Figure 8a). The line in Figure 8a shows the deviation from the linear law. It should be also noted that for other temperature ranges during cooling, the additional nonlinear dependences of the ∆L/L(T) are formed (Figure 8a). This indicates the presence of certain structural instabilities that do not significantly affect the dielectric properties, but apparently change the linear dimensions of the experimental sample. The dependence α(T) is in a good correlation with ∆L/L(T); they have extremes in the similar temperature ranges (Figure 8a).
With the introduction of the large-sized La and Pr ions (Figure 8b and Figure 9a), thermophysical properties do not undergo any significant changes over the entire investigated temperature range. This may be due to the small distortions of the unit cell in the PsC phase and the proximity of its configuration to the cubic syngony, and the change in the temperature only facilitates the transition to the purely cubic phase.
For the SSs containing Tb, Dy, and Ho (Figure 9b and Figure 10a,b), an almost monotonic increase in the ∆L/L dependence is characteristic with a change in the slope of the curves in the vicinity of the temperatures of the beginning and end of the plateau-like portion of the ε′/ε0(T) dependence. This occurs during cooling especially clearly in the SS with Dy (Figure 10a).

3.5. Microhardness

The results of studying the microhardness (H) of the ceramics by the comparative sclerometry method [52] and Young’s modulus (E) determined from the load-displacement curves [53] are presented in Table 2. To reduce the standard error of the data, the microhardness values obtained at various loads, were averaged over several measurements. The results made it possible to estimate the stress intensity factor for mode I, Kic, which is a criterion for the crack resistance of the material. The crack resistance Kic was determined in accordance with the model of Anstis et al. [54,55] by the formula:
K i c = 0.016 E H 1 2 P C 3 2 ,  
where P—applied load, E—Young’s modulus, H—microhardness, and c—average distance from the center of the indentation to the end of the crack.
In general, all the considered UHP [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3 ceramics have high hardness and Young’s modulus and, as a consequence, good crack resistance for the ceramic materials (Table 2). However, the series contains samples of the SS modified with La and Pr; they have a pseudocubic structure, and their hardness is H = 6.11 ± 0.62 GPa (La) and H = 8.42 ± 1.33 GPa (Pr). Such a significant change in the microhardness of the ceramics is probably associated with a change in the symmetry of the perovskite cell and an increase in the syngony from monoclinic (lower category) to pseudocubic (higher category) for the [(Na0.90K0.05Cd0.05)0.95La0.05]NbO3 and [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3 SS (Table 1 and Table 2).

3.6. Photoluminescence

The initial and modified with REE ceramics were transparent enough during UHP preparation (Figure 11). We decided to measure not only mechanical characteristics, but also optical (luminescent) characteristics of the pure and doped with 5 mol% of the REE (La, Pr, Tb, Dy, Ho) Na0.90K0.05Cd0.05NbO3 ceramics.
Figure 12 shows PL spectra of the pure and modified with the REE (La3+, Pr3+, Tb3+, Dy3+, Ho3+) Na0.90K0.05Cd0.05NbO3 ceramic SSs. All PL spectra have two characteristic regions: from 380 to 575 nm and from 575 to 800 nm.
In the first region, a broad luminescent halo is observed with a maximum at about 500 nm. In the second region, several maxima of different intensities and widths are observed at ~610, 710, and 775 nm. In this case, the [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3 sample with a pseudocubic structure has the maximal PL.
In the unmodified Na0.90K0.05Cd0.05NbO3 ceramic sample, a very weak luminescent halo is observed with a maximum at ~500 nm (Figure 12 (curve 1)). This halo can be associated with the radiative recombination between the Nb5+ ion and the surrounding O2− oxygen ions [56]. In addition, depending on the composition of the ceramics, the maximum of the luminescent halo for NaNbO3 can shift from 480 to 540 nm, and the luminescence intensity decreases by half when the temperature changes from 90 to 50 K [57]. The KNbO3 spectrum exhibits two luminescence maxima corresponding to its two polymorphic modifications: at 420 and 530 nm. In this case, the temperature quenching of the luminescence is observed at 70 and 30 K [57]. Thus, the low luminescence intensity of the Na0.90K0.05Cd0.05NbO3 ceramic matrix under normal conditions (T = 300 K) is apparently due to the presence of the temperature quenching of the radiative recombination of the Nb-O pair. In addition, it is possible that the energy transfers between different Nb-O pairs, surrounded by the metal cations with different ionic radii (Na, K, Cd—98, 133, 99 pm, respectively [58]), and then scatters on the vibrations of the crystal lattice.
Upon the modification of the Na0.90K0.05Cd0.05NbO3 SS with the REE, the shape of the spectrum changes and the intensity of the luminescence intensity distributes between the matrix itself and the characteristic intraconfigurational 4fn-4fn transitions of the REE (Figure 12 (2–6)).
Thus, the characteristic radiative transitions for Pr3+-3P0,1,2-3H4, 1D2-3H4, and 3P0-3F2 are located in the regions of 450–500, 620, and 660 nm, and the intensity of the maxima at 450–500 and 660 nm is minimal [59]. The typical radiative transitions for Dy3+ are 4F9/2-6HJ (J = 15/2, 13/2, and 11/2) at 476, 573, and 667 nm [60,61]; for Tb3+-5D3-7F4,5, and 5D4-7FJ (J = 6,5,4, and 3) at 442, 492, 548, 588, and 623 nm [61]; for Ho3+-5F4-5I8, 5S2-5I8, and 5F5-5I8 at 541, 551, and 653 nm [62].
The La3+ cation apparently lacks the 4fn-4fn transitions, since its electronic configuration is 4f05d16s2. Due to the fact that the excitation of the spectrum (λex = 325 nm) is performed in the fundamental absorption region (3.41 eV for NaNbO3 [63]), initially the spectrum of the ceramic matrix is excited; that is, the Nb-O pairs absorb and then radiate at 500 nm. Due to the charge transfer in the Pr3+-Men+ (Men+ = Ti4+, Nb5+, Ta5+, etc.) pair, the population of the 1D2 state [63] is possible, from which the electrons pass into the basic 3H4 state. In this case, the strong luminescence is observed in the region of ~610 nm, which is clearly seen in the spectrum of the [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3 ceramics (Figure 12 (6)). However, Figure 12 (2, 3) shows that there are no effective radiative recombination of the intraconfigurational 4fn-4fn transitions in the PL spectra of the Na0.90K0.05Cd0.05NbO3 samples modified with Tb3+ and Ho3+.
This can be explained by the fact that during the synthesis of the ceramics, the valence of Tb can change from 3+ to 4+, and with an increase in the annealing temperature to 1473 K, the number of Tb4+ ions also increases [59]. Such Tb4+ ions act as the quenching centers and not only reduce the amount of the Tb3+ ions, but also absorb the radiative recombination of the 5D3,4-7FJ transitions [59]. Thus, it can be assumed that during the synthesis of the [(Na0.90K0.05Cd0.05)0.95Tb0.05]NbO3 ceramics, the number of the Tb4+ cations increases, which acts as a luminescence quencher in the visible region. The Ho3+ ions are usually used as an activator in the up-conversion processes, where together with a sensitizer, usually the Yb3+ ions [64], the luminescence intensity is increased in the visible region excited by IR radiation. Thus, in the absence of a sensitizer, the probability of the radiative transitions in the visible region for Ho is small, and PL spectra of the [(Na0.90K0.05Cd0.05)0.95Ho0.05]NbO3 ceramics prove it (Figure 12 (2)).
On the other hand, the Dy3+ ions in the Na0.90K0.05Cd0.05NbO3 matrix enhance the luminescence of the crystalline matrix. In this case, the PL intensity is five times higher than in the unmodified ceramics (Figure 12 (1, 4)). Since the formation of the Dy2+ electron trap is possible only when exposed to the ionizing radiation, the luminescence of which is observed in the IR region at 1450 nm [65], the energy converses between the Dy3+ cations and the matrix. Moreover, the maximum luminescence intensity in the spectrum of [(Na0.90K0.05Cd0.05)0.95Dy0.05]NbO3 ceramics is in the range of 480–540 nm (Figure 12 (4)). It should also be noted that an increase in the symmetry of the crystal structure of the samples under study can lead to an increase in the probability of the radiative transitions and energy transfer between the matrix and the REE cations in the visible wavelength range. This can be seen in the example of the [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3 and [(Na0.90K0.05Cd0.05)0.95La0.05]NbO3 SSs, for which an increase in the symmetry of the structure leads to an increase in the luminescence intensity (Table 2 and Figure 8 (5, 6)).

4. Conclusions

Ceramic solid solutions (SSs) with the composition Na0.90K0.05Cd0.05NbO3 and [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3 (where REE = La, Pr, Tb, Dy, Ho) were synthesized for the first time by uniaxial hot pressing (UHP). XRD analysis showed changes in the structural characteristics of the SS Na0.90K0.05Cd0.05NbO3 upon its modification with REEs La, Pr, Tb, Dy, and Ho. XRD also proved the absence of impurities. The volume of the perovskite unit cell slightly increases upon the introduction of large-size ions La3+ and Pr3+, which indicates their incorporation into the crystal lattice of the initial solid solution. It was shown that the REEs Tb, Dy, and Ho do not completely enter the crystal lattice of the initial solid solution, but form impurity compounds LnNbO4 of the monoclinic fergusonite type.
It was found that UHP ceramics have a homogeneous fine-grained structure with increased mechanical characteristics: hardness, Young’s modulus, and high crack resistance.
A study of the dielectric and thermophysical properties in a wide temperature range revealed the formation in the initial solid solution Na0.90K0.05Cd0.05NbO3 of two peaks of the dielectric constant. They correspond to the transition from orthorhombic to tetragonal (TO-T) and from tetragonal to cubic (TT-C) phase. The introduction of large-sized REEs shifted TO-T and TT-C to the low-temperature region. However, introduction of middle-and small-sized REEs shifts T→C transition to a low-temperature and O→T transition to a high-temperature region; they merge and form a quasi-plateau region. This may be caused by the different incorporation mechanisms of REEs to the crystal lattice of investigated SSs.
The dispersion of the permittivity is observed in the paraelectric region of both original and modified SSs. This may be due to the redox processes that exist due to the variable valence of niobium. Redox processes cause the appearance of oxygen vacancies, and as a result, increase electrical conductivity and its contribution to the permittivity.
Both the real and imaginary parts of the relative permittivity are blurred with the introduction to SS of all studied modifiers. The greatest influence is exerted by medium- and small-sized REE. This might be explained by the following: they partially incorporate into the structure and form non-ferroelectric ballast phases.
Anomalies of thermophysical characteristics correspond to the above-described changes in dielectric properties in the entire studied temperature range: in the vicinity of phase transitions, the thermophysical characteristics are extreme. The thermophysical characteristics experience changes in those intervals where the dielectric properties do not exhibit anomalous behavior. This indicates the existence of small structural instabilities. However, confirmation requires additional XRD studies.
Very weak luminescence intensity from the electron-hole pair Nb4+-O- was observed in the green region with a maximum at ~500 nm for the ceramic SS Na0.90K0.05Cd0.05NbO3. At the same time, in the Na0.90K0.05Cd0.05NbO3 SS and SSs modified with Ho and Tb a low intense luminescence of the ceramic matrix is observed in comparison with the intensity of luminescence of the matrix in the Na0.90K0.05Cd0.05NbO3 samples modified with La and Dy. The intensity is low due to the energy dissipation on the crystal lattice vibrations and the influence of the REE cations on the charge transfer between different NbO groups. Thus, the modification by Tb and Ho does not lead to an increase in the luminescence of the Na0.90K0.05Cd0.05NbO3 ceramics in the visible region, in contrast to the modification by Dy3+. In the modification by Dy3+ case the luminescence intensity in the visible region increases by a factor of five. An increase in the symmetry of the [(Na0.90K0.05Cd0.05)0.95La(Pr)0.05]NbO3 SS leads to a more efficient radiative recombination between Nb-O. This is apparently due to the efficient transfer of the energy between the matrix and REE, which, in particular, leads to very intense luminescence due to the radiative 1D2-3H4 transition of the Pr3+ ion in the [(Na0.90K0.05Cd0.05)0.95Pr3+0.05]NbO3 ceramics near ~610 nm. The study of optical characteristics of [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3 SSs synthesized in this study widens their potential application as luminescent materials.

Author Contributions

A.K.: Conceptualization, Methodology, Investigation, Supervision, Writing—Original Draft, Visualization, Investigation, Validation, Methodology, Data Curation. M.P.: Conceptualization, Writing—Original Draft. L.S.: Investigation, Writing—Review and Editing. A.N.: Investigation, Visualization. O.S.: Conceptualization, Writing—Original Draft, Writing—Review and Editing, Visualization, Investigation, Validation, Methodology, Data Curation. M.S.: Writing—Original Draft, Investigation, Validation, Methodology, Data Curation. N.S.: Writing—Original Draft, Investigation, Validation. S.K.: Investigation, Validation, Writing—Review and Editing. A.R.: Visualization, Data Curation, Validation. D.R.: Investigation, Visualization. L.R.: Conceptualization, Writing—Original Draft, Supervision, Writing—Review and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

The study was carried out with the financial support of the Ministry of Science and Higher Education of the Russian Federation, topicFENW-2023-0010.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data required to reproduce these findings are available from the corresponding author on a reasonable request.

Acknowledgments

The equipment of the Center of Research Institute of Physics SFedU, “High-Tech” SFedU was used in this study. We thank the center for the given opportunity.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Von Hippel, A. Ferroelectricity, domain structure, and phase transitions of barium titanate. Rev. Modern Phys. 1950, 22, 221–237. [Google Scholar] [CrossRef]
  2. Merz, W.J. Domain properties in BaTiO3. Phys. Rev. 1952, 88, 421–422. [Google Scholar] [CrossRef]
  3. Landauer, R. Electrostatic considerations in BaTiO3 domain formation during polarization reversal. J. Appl. Phys. 1957, 28, 227–234. [Google Scholar] [CrossRef]
  4. Last, J.T. Infrared-absorption studies on barium titanate and related materials. Phys. Rev. 1957, 105, 1740–1750. [Google Scholar] [CrossRef]
  5. Wentz, J.L.; Kennedy, L.Z. Primary pyroelectric effect in the PZT 95/5 ceramic. J. Appl. Phys. 1964, 35, 1767–1770. [Google Scholar] [CrossRef]
  6. Linde, R.K. Depolarization of ferroelectrics at high strain rates. J. Appl. Phys. 1967, 38, 4839–4842. [Google Scholar] [CrossRef]
  7. Carruthers, J.R.; Grasso, M. Phase equilibria relations in the ternary system BaO-SrO-Nb2O5. J. Electrochem. Soc. 1970, 117, 1426. [Google Scholar] [CrossRef]
  8. Itoh, Y.; Lwasaki, H. Tungsten bronze field in the pseudo-ternary system NaNbO5-BaTiO3-BaNb2O6. Mat. Res. Bull. 1972, 7, 663–672. [Google Scholar] [CrossRef]
  9. Ogawa, T.; Wakino, K. The temperature property of frequency constant of Pb(Sn1/2Sb1/2)O3-PbTiO3-PbZrO3 piezoelectric ceramics. J. Jpn. Soc. Powder Powder Metall. 1977, 24, 123–127. [Google Scholar] [CrossRef]
  10. Gonnard, P.; Troccaz, M. Dopant Distribution between A and B Sites in the PZT Ceramics of Type ABO3. J. Sol. St. Chem. 1978, 23, 321–326. [Google Scholar] [CrossRef]
  11. Tsubouchi, N.; Takahashi, M.; Chino, T.; Akashi, T.; Tokyo-to, Japan, Tokyo, Assignors to Nippon Electric Company, Limited. Piezoelectric Ceramic. US Patent No. 3595795, 1971. [Google Scholar]
  12. Nippon Electric Company, Limited, Tokyo, Japan. Piezoelectric Ceramic Material. France Patent No. FR1580831A, 28 December 1969. [Google Scholar]
  13. Murty, S.N.; Ramana, K.V.M.; Umakantham, K.; Bhanumathi, A. Modified (NaK)NbO3 ceramics for transducer applications. Ferroelectrics 1990, 102, 243–247. [Google Scholar] [CrossRef]
  14. Murty, S.N. Piezoelectric (NaK)NbO3 ceramics for ultrasonic delay line applications. Ferroelectrics 1994, 154, 171–176. [Google Scholar] [CrossRef]
  15. Shan, Y.; Sinozaki, N.; Nakamura, T. Preparation and characterizations of new perovskite oxides (LaxNa1−3x−yLiy2x)NbO3 (0.0 ≤ x and y ≤ 0.2). Sol. St. Ion. 1998, 108, 403–406. [Google Scholar] [CrossRef]
  16. Wang, X.; Helmersson, U.; Olafsson, S.; Rudner, S.; Wernlund, L.D.; Gevorgian, S. Growth and field dependent dielectric properties of epitaxial Na0.5K0.5NbO3 thin films. Appl. Phys. Lett. 1998, 73, 927–929. [Google Scholar] [CrossRef]
  17. EU. Commision Directive 2013/28/EU of 17 May 2013 amending annex II to directive 2000/53/EC of the European Parliament and of the Council on end-of-life vehicles. Off. J. Eur. Union 2013, L135, 14–18. [Google Scholar]
  18. Zhang, S.T.; Kounga, A.B.; Aulbach, E.; Ehrenberg, H.; Rödel, J. Giant strain in lead-free piezoceramics Bi0.5Na0.5TiO3-BaTiO3-K0.5Na0.5NbO3 system. Appl. Phys. Lett. 2007, 91, 2–5. [Google Scholar] [CrossRef]
  19. Rödel, J.; Jo, W.; Seifert, K.T.P.; Anton, E.M.; Granzow, T.; Damjanovic, D. Perspective on the development of lead-free piezoceramics. J. Amer. Ceram. Soc. 2009, 92, 1153–1177. [Google Scholar] [CrossRef]
  20. Matsubara, M.; Yamaguchi, T.; Wataru, S.; Koichi, K.; Yogo, T.; Hirano, S. Processing and Piezoelectric Properties of Lead-Free (K, Na)(Nb, Ta)O3 Ceramics. J. Am. Ceram. Soc. 2005, 88, 1190–1196. [Google Scholar] [CrossRef]
  21. Verbenko, I.A.; Razumovskaya, O.N.; Shilkina, L.A.; Reznichenko, L.A.; Andryushin, K.P. Production and dielectric properties of lead-free ceramics with the formula [(Na0.5K0.5)1-xLix](Nb1-y-zTaySbz)O3. Inorg. Mater. 2009, 45, 702–708. [Google Scholar] [CrossRef]
  22. Saito, Y.; Takao, H.; Tani, T.; Nonoyama, T.; Takatori, K.; Homma, T.; Nagaya, T.; Nakamura, M. Lead-free piezoceramics. Nature 2004, 432, 84–87. [Google Scholar] [CrossRef]
  23. Li, P.; Zhai, J.; Shen, B.; Zhang, S.; Li, X.; Zhu, F.; Zhang, X. Ultrahigh piezoelectric properties in textured (K, Na)NbO3-based lead-free ceramics. Adv. Mater. 2018, 30, 1705171. [Google Scholar] [CrossRef]
  24. Wu, J.; Xiao, D.; Zhu, J. Potassium-sodium niobate lead-free piezoelectric materials: Past, present, and future of phase boundaries. Chem. Rev. 2015, 115, 2559–2595. [Google Scholar] [CrossRef]
  25. Jo, W.; Dittmer, R.; Acosta, M.; Zang, J.; Groh, C.; Sapper, E.; Wang, K.; Rödel, J. Giant electric-field-induced strains in lead-free ceramics for actuator applications—Status and perspective. J. Electroceram. 2012, 29, 71–93. [Google Scholar] [CrossRef]
  26. Li, W.; Hao, J.; Li, W.; Du, J.; Fu, P.; Sun, W.; Chen, C.; Xu, Z.; Chu, R. Electrical properties and luminescence properties of 0.96(K0.48Na0.52)(Nb0.95Sb0.05)–0.04Bi0.5(Na0.82K0.18)0.5ZrO3-xSm lead-free ceramics. J. Adv. Ceram. 2020, 9, 72–82. [Google Scholar] [CrossRef]
  27. Andryushin, K.P.; Shilkina, L.A.; Andryushina, I.N.; Moysa, M.O.; Rudsky, D.I.; Reznichenko, L.A. Crystal structure, polarization properties and reverse nonlinearity of solid solutions of the KNN-PZT system in a wide range of external influences. Ceram. Intern. 2021, 47, 138–148. [Google Scholar] [CrossRef]
  28. Nagaenko, A.V.; Chang, S.H.; Andryushin, K.P.; Shilkina, L.A.; Mazuritskiy, M.I.; Andryushina, I.N.; Glazunova, E.V.; Pavelko, A.A.; Trusov, Y.A.; Verbenko, I.A.; et al. Multi-element ferroactive materials based on KNN-PZT compositions with fundamentally different physical properties. Heliyon 2020, 6, e03497. [Google Scholar] [CrossRef] [PubMed]
  29. Rödel, J.; Webber, K.G.; Dittmer, R.; Jo, W.; Kimura, M.; Damjanovic, D. Transferring lead-free piezoelectric ceramics into application. J. Europ. Ceram. Soc. 2015, 35, 1659–1681. [Google Scholar] [CrossRef]
  30. Andryushin, K.; Shilkina, L.; Andryushina, I.; Nagaenko, A.; Moysa, M.; Dudkina, S.; Reznichenko, L. Features of the Structure and Electrophysical Properties of Solid Solutions of the System (1-x-y) NaNbO3-xKNbO3-yCd0.5NbO3. Materials 2021, 14, 4009. [Google Scholar] [CrossRef]
  31. Isupov, V.A. Physical phenomena in ferroelectric complex perovskites. Bull. Acad. Sci. U.S.S.R. Phys. Ser. 1982, 47, 136–141. [Google Scholar]
  32. Von der Mühll, R.; Sadel, A.; Hagenmuller, P. Structure cristallinea 295 K de la phase ferroelectrique Li0,02Na0,98NbO3. J. Sol. St. Chem. 1984, 51, 176–182. [Google Scholar] [CrossRef]
  33. Isupov, V.A. Crumbling -type phase transitions. Cryst. Rep. 1959, 4, 603–608. (In Russian) [Google Scholar]
  34. Megaw, H.D. The Seven Phases of Sodium Niobate. Ferroelectrics 1974, 7, 87–89. [Google Scholar] [CrossRef]
  35. Chen, J.; Feng, D. In situ TEM studies of para-ferro phase transitions in NaNbO3. Phys. Stat. Sol. A 1988, 109, 427–434. [Google Scholar] [CrossRef]
  36. Reznichenko, L.A.; Shilkina, L.A. Morphotropic Phase in the NaNbO3-LiNbO3 Solid-Solution System. Bull. Acad. Sci. U.S.S.R. Phys. Ser. 1975, 39, 1118–1121. [Google Scholar]
  37. Reznichenko, L.A.; Shilkina, L.A.; Razumovskaya, O.N.; Dudkina, S.I.; Gagarina, E.S.; Borodin, A.V. Dielectric and piezoelectric properties of NaNbO3-based solid solutions. Inorg. Mater. 2003, 39, 139–151. [Google Scholar] [CrossRef]
  38. Fesenko, E.G. The Perovskite Family and Ferroelectricity; (In Russian: Semejstvo Perovskita I Segnetoehlektrichestvo); Atomizdat: Moscow, Russia, 1972. [Google Scholar]
  39. JCPDS. Powder Diffraction File; Data Card. Inorganic Section. Set, card; JCPDS: Swarthmore, PA, USA, 1948. [Google Scholar]
  40. Urusov, V.S. Theory of Isomorphic Miscibility; Science: Moscow, Russia, 1977; (In Russian: Teoriya izomorfnoy smesimosti). [Google Scholar]
  41. Dantsiger, A.Y.; Reznitchenko, L.A.; Dudkina, S.I.; Razumovskaya, O.N.; Shilkina, L.A. Correlation between the microstructure of ferroelectric ceramics and their chemical and phase composition, the degree of perfection of the crystal structure and the preparation conditions. Ferroelectrics 1998, 214, 255–259. [Google Scholar] [CrossRef]
  42. Rao, C.N.R.; Gopalakrishnan, J. New Directions in Solid State Chemistry; Cambridge University Press: Cambridge, UK, 1986. [Google Scholar]
  43. Hedvall, J.A. Über die physikalisch-chemischen Prozesse beim Zusammenbacken von angeschmolzenen Pulvern. Z. Phys. Chem. 1926, 123U, 33–85. [Google Scholar] [CrossRef]
  44. Liu, L.; Huang, Y.; Li, Y.; Fang, L.; Dammak, H.; Fan, H.; Thi, M.P. Orthorhombic to tetragonal structural phase transition in Na0.5K0.5NbO3-based ceramics. Mater. Lett. 2012, 68, 300–302. [Google Scholar] [CrossRef]
  45. Du, H.; Huang, Y.H.; Tang, H.; Qin, H.; Feng, W. Dielectric and piezoelectric properties of SrZrO3 -modified (K0.45Na0.51Li0.04)(Nb0.90Ta0.04Sb0.06)O3 lead-free piezoceramics. Mater. Lett. 2013, 106, 141–144. [Google Scholar] [CrossRef]
  46. Shi, W.; Feng, Y.; Lu, T.; Lu, Y.; Shen, J.; Xue, J.; Du, J.; Fu, P.; Hao, J.; Li, W. Photoluminescence and impedance properties of rare-earth doped (K0.5Na0.5)NbO3 lead-free ceramics. J. Mater. Sci. Mater. Electron. 2019, 30, 9–16. [Google Scholar] [CrossRef]
  47. Egerton, L.; Dillon, D.M. Piezoelectric and Dielectric Properties of Ceramics in the System Potassium—Sodium Niobate. J. Amer. Ceram. Soc. 1959, 42, 438–442. [Google Scholar] [CrossRef]
  48. Zhang, N.; Lv, X.; Zhang, X.-X.; Lyu, X.; Yang, S.-W.; Wu, J. Low-temperature dielectric relaxation associated with NbO6 octahedron distortion in antimony modified potassium sodium niobate ceramics. J. Mater. Sci. Technol. 2022, 115, 189–198. [Google Scholar] [CrossRef]
  49. Uchino, K.; Nomura, S. Critical exponents of the dielectric constants in diffused-phase-transition crystals. Ferroelectrics 1982, 44, 55–61. [Google Scholar] [CrossRef]
  50. Liu, L.; Knapp, M.; Ehrenberg, H.; Fang, L.A.; Fan, H.; Schmitt, L.; Fuess, H.; Hoelzel, M.; Dammak, H.; Thi, M.P.; et al. Average vs. local structure and composition-property phase diagram of K0.5Na0.5NbO3-Bi½Na½TiO3 system. J. Europ. Ceram. Soc. 2017, 37, 1387–1399. [Google Scholar] [CrossRef]
  51. Zhai, Y.; Du, J.; Chen, C.; Hao, J.; Fu, P.; Li, W.; Xu, Z. Temperature stability and electrical properties of Tm2O3 doped KNN-based ceramics. J. Mater. Sci. Mater. Electron. 2019, 30, 4716–4725. [Google Scholar] [CrossRef]
  52. Oliver, W.; Pharr, G.M. Measurement of hardness and elastic modulus by instrumented indentation: Advances in understanding and refinements to methodology. J. Mater. Res. 2004, 19, 3. [Google Scholar] [CrossRef]
  53. Maslenikov, I.I.; Reshetov, V.N.; Useinov, A.S. Mapping the elastic modulus of a surface with a NanoScan 3D scanning microscope. Instrum. Experim. Tech. 2015, 58, 711–717. [Google Scholar] [CrossRef]
  54. Anstis, G.R.; Chantikul, P.; Lawn, B.R.; Marshall, D. A Critical Evaluation of Indentation Techniques for Measuring Fracture Toughness: I, Direct Crack Measurements. Transformation 1981, 46, 533–538. [Google Scholar] [CrossRef]
  55. Chantikul, P.; Anstis, G.R.; Lawn, B.R.; Marshall, D. A Critical Evaluation of Indentation Techniques for Measuring Fracture Toughness: II, Strength Method. J. Amer. Ceram. Soc. 1981, 64, 539–543. [Google Scholar] [CrossRef]
  56. Blasse, G.; Bril, A. Luminescence phenomena in compounds with fergusonite structure. J. Luminesc. 1970, 3, 109–131. [Google Scholar] [CrossRef]
  57. Blasse, G.; De Haart, L. The nature of the luminescence of niobates MNbO3 (M = Li, Na, K). Mater. Chem. Phys. 1986, 14, 481–484. [Google Scholar] [CrossRef]
  58. Shaskolskaya, M.P. Crystallography: Textbook. Manual for Technical Colleges, 2nd ed.; Higher School: Moscow, Russia, 1984; (In Russian: Kristallografiya: Ucheb. posobiye dlya vtuzov. -2-ye izd., pererab. i dop.). [Google Scholar]
  59. Verma, R.K.; Kumar, K.; Rai, S. Inter-conversion of Tb3+ and Tb4+ states and its fluorescence properties in MO-Al2O3: Tb (M = Mg, Ca, Sr, Ba) phosphor materials. Sol. St. Sci. 2010, 12, 1146–1151. [Google Scholar] [CrossRef]
  60. Wen, R.; Zhou, L.; Zou, X.; Luo, L.; Jiang, N.; Zheng, Q.; Liao, J.; Xu, C.; Lin, D. Phase transition, electrical and luminescent properties of Dy-doped K0.5Na0.5NbO3-based lead-free ceramics. J. Mater. Sci: Mater. Electron. 2015, 26, 8341–8349. [Google Scholar] [CrossRef]
  61. Ramachari, D.; Moorthy, L.R.; Jayasankar, C.K. Energy transfer and photoluminescence properties of Dy3+/Tb3+ co-doped oxyfluorosilicate glass-ceramics for solid-state white lighting. Ceram. Intern. 2014, 40, 11115–11121. [Google Scholar] [CrossRef]
  62. Zhang, Y.; Liu, J.; Sun, H.; Peng, D.; Li, R.; Bulin, C.; Wang, X.; Zhang, Q.; Hao, X. Reversible luminescence modulation of Ho-doped K0.5Na0.5NbO3 piezoelectrics with high luminescence contrast. J. Am. Ceram. Soc. 2018, 101, 2305–2312. [Google Scholar] [CrossRef]
  63. Yang, X.; Zhao, L.; Chen, W.; Liu, Z.; Fan, X.; Tian, S.; Xu, X.; Qiu, J.; Yu, X. Low-temperature red long-persistent luminescence of Pr3+ doped NaNbO3 with a perovskite structure. J. Luminesc. 2019, 208, 290–295. [Google Scholar] [CrossRef]
  64. Kir’yanov, A.V.; Aboites, V.; Belovolov, A.M.; Damzen, M.; Minassian, A.; Timoshechkin, M.; Belovolov, M. Visible-to-near-IR luminescence at stepwise up-conversion in Yb, Ho: GGG under IR diode pumping. J. Lumines. 2003, 102–103, 715–721. [Google Scholar] [CrossRef]
  65. Schuyt, J.J.; Williams, G.V.M. Photoluminescence of Dy3+ and Dy2+ in NaMgF3:Dy: A potential infrared radiophotoluminescence dosimeter. Radiat. Meas. 2020, 134, 106326. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of the studied SSs: (1)—the initial and modified with (2) La; (3) Pr; (4) Tb; (5) Dy; (6) Ho. Indices (hkl)c correspond to the perovskite axes. The arrows mark the peaks of the impurity phases: TbNbO4 (4) (JCPDS 23-1417), DyNbO4 (5) (JCPDS 22-1094), and HoNbO4 (6) (JCPDS 23-1104). The angular interval is marked by vertical segments. (*) asterisk marks a very weak superstructural peak indicating the doubling of the b axis. (b) The diffraction peaks (200) and (020) are enlarged along the 2θ axis.
Figure 1. (a) XRD patterns of the studied SSs: (1)—the initial and modified with (2) La; (3) Pr; (4) Tb; (5) Dy; (6) Ho. Indices (hkl)c correspond to the perovskite axes. The arrows mark the peaks of the impurity phases: TbNbO4 (4) (JCPDS 23-1417), DyNbO4 (5) (JCPDS 22-1094), and HoNbO4 (6) (JCPDS 23-1104). The angular interval is marked by vertical segments. (*) asterisk marks a very weak superstructural peak indicating the doubling of the b axis. (b) The diffraction peaks (200) and (020) are enlarged along the 2θ axis.
Ceramics 06 00021 g001
Figure 2. SEM images of transverse cleavage of ceramic SSs: (a) of the initial one and modified with (b) La; (c) Pr; (d) Tb; (e) Dy; (f) Ho. Insets demonstrate differential curves of ceramic grain size distribution in corresponding composition of SS.
Figure 2. SEM images of transverse cleavage of ceramic SSs: (a) of the initial one and modified with (b) La; (c) Pr; (d) Tb; (e) Dy; (f) Ho. Insets demonstrate differential curves of ceramic grain size distribution in corresponding composition of SS.
Ceramics 06 00021 g002
Figure 3. Column (1): dependence of ε′/ε0; column (2): dependence of ε″/ε0 on the temperature at different frequencies (f) of the alternating electric field obtained in the cooling mode of the samples, in temperature range of T = (300–900) K: (a) Na0.90K0.05Cd0.05NbO3; (b) [(Na0.90K0.05Cd0.05)0.95La0.05]NbO3; (c) [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3. The inset in a(1) shows dependences (ε′/ε0)−1(T) to demonstrate the empirical Curie-Weiss law. The insets in b(1) and c(1) show the dispersion and blurring of ε′/ε0 at the instant of the phase transition. The arrow shows the direction of f increase. Frequencies are denoted as follows: Ceramics 06 00021 i001 8∙102 Hz; Ceramics 06 00021 i002 1.0 104 Hz; Ceramics 06 00021 i003 1.0∙105 Hz; Ceramics 06 00021 i004 5.2∙105 Hz; Ceramics 06 00021 i005 2.5∙106 Hz; Ceramics 06 00021 i006 7.2∙106 Hz; Ceramics 06 00021 i007 1.2∙107 Hz.
Figure 3. Column (1): dependence of ε′/ε0; column (2): dependence of ε″/ε0 on the temperature at different frequencies (f) of the alternating electric field obtained in the cooling mode of the samples, in temperature range of T = (300–900) K: (a) Na0.90K0.05Cd0.05NbO3; (b) [(Na0.90K0.05Cd0.05)0.95La0.05]NbO3; (c) [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3. The inset in a(1) shows dependences (ε′/ε0)−1(T) to demonstrate the empirical Curie-Weiss law. The insets in b(1) and c(1) show the dispersion and blurring of ε′/ε0 at the instant of the phase transition. The arrow shows the direction of f increase. Frequencies are denoted as follows: Ceramics 06 00021 i001 8∙102 Hz; Ceramics 06 00021 i002 1.0 104 Hz; Ceramics 06 00021 i003 1.0∙105 Hz; Ceramics 06 00021 i004 5.2∙105 Hz; Ceramics 06 00021 i005 2.5∙106 Hz; Ceramics 06 00021 i006 7.2∙106 Hz; Ceramics 06 00021 i007 1.2∙107 Hz.
Ceramics 06 00021 g003aCeramics 06 00021 g003b
Figure 4. Column (1): dependence of ε′/ε0; column (2): dependence of ε″/ε0 on the temperature at different frequencies (f) of the alternating electric field obtained in the cooling mode of the samples, in temperature range of T = (300–900) K: (a) [(Na0.90K0.05Cd0.05)0.95Tb0.05]NbO3; (b) [(Na0.90K0.05Cd0.05)0.95Dy0.05]NbO3; (c) [(Na0.90K0.05Cd0.05)0.95Ho0.05]NbO3. The insets in column (1) show the dispersion and blurring of ε′/ε0 at the instant of the phase transition. The arrow shows the direction of f increase. Frequencies are denoted as follows: Ceramics 06 00021 i0018 ∙102 Hz; Ceramics 06 00021 i002 1.0 104 Hz; Ceramics 06 00021 i003 1.0∙105 Hz; Ceramics 06 00021 i004 5.2∙105 Hz; Ceramics 06 00021 i005 2.5∙106 Hz; Ceramics 06 00021 i006 7.2∙106 Hz; Ceramics 06 00021 i007 1.2∙107 Hz.
Figure 4. Column (1): dependence of ε′/ε0; column (2): dependence of ε″/ε0 on the temperature at different frequencies (f) of the alternating electric field obtained in the cooling mode of the samples, in temperature range of T = (300–900) K: (a) [(Na0.90K0.05Cd0.05)0.95Tb0.05]NbO3; (b) [(Na0.90K0.05Cd0.05)0.95Dy0.05]NbO3; (c) [(Na0.90K0.05Cd0.05)0.95Ho0.05]NbO3. The insets in column (1) show the dispersion and blurring of ε′/ε0 at the instant of the phase transition. The arrow shows the direction of f increase. Frequencies are denoted as follows: Ceramics 06 00021 i0018 ∙102 Hz; Ceramics 06 00021 i002 1.0 104 Hz; Ceramics 06 00021 i003 1.0∙105 Hz; Ceramics 06 00021 i004 5.2∙105 Hz; Ceramics 06 00021 i005 2.5∙106 Hz; Ceramics 06 00021 i006 7.2∙106 Hz; Ceramics 06 00021 i007 1.2∙107 Hz.
Ceramics 06 00021 g004aCeramics 06 00021 g004b
Figure 5. Column (1): dependence ε′/ε0, column (2): dependence ε″/ε0 on the temperature at different frequencies (f) of the alternating electric field obtained in the cooling mode of the samples, in temperature range of T = (10–325) K: (a) Na0.90K0.05Cd0.05NbO3; (b) [(Na0.90K0.05Cd0.05)0.95La0.05]NbO3; (c) [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3. The insets in column (1) show the dispersion and blurring of ε’/ε0 at the instant of the phase transition. The arrow shows the direction of f increase. Frequencies are denoted as follows: Ceramics 06 00021 i008 96 Hz; Ceramics 06 00021 i001 8∙102 Hz; Ceramics 06 00021 i002 1.0 104 Hz; Ceramics 06 00021 i003 1.0∙105 Hz; Ceramics 06 00021 i004 2.5∙105 Hz; Ceramics 06 00021 i005 7.2∙106 Hz; Ceramics 06 00021 i006 7.2∙106 Hz; Ceramics 06 00021 i007 1.2∙107 Hz.
Figure 5. Column (1): dependence ε′/ε0, column (2): dependence ε″/ε0 on the temperature at different frequencies (f) of the alternating electric field obtained in the cooling mode of the samples, in temperature range of T = (10–325) K: (a) Na0.90K0.05Cd0.05NbO3; (b) [(Na0.90K0.05Cd0.05)0.95La0.05]NbO3; (c) [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3. The insets in column (1) show the dispersion and blurring of ε’/ε0 at the instant of the phase transition. The arrow shows the direction of f increase. Frequencies are denoted as follows: Ceramics 06 00021 i008 96 Hz; Ceramics 06 00021 i001 8∙102 Hz; Ceramics 06 00021 i002 1.0 104 Hz; Ceramics 06 00021 i003 1.0∙105 Hz; Ceramics 06 00021 i004 2.5∙105 Hz; Ceramics 06 00021 i005 7.2∙106 Hz; Ceramics 06 00021 i006 7.2∙106 Hz; Ceramics 06 00021 i007 1.2∙107 Hz.
Ceramics 06 00021 g005
Figure 6. Column (1): dependence ε′/ε0, column (2): dependence ε″/ε0 on the temperature at different frequencies of the alternating electric field obtained in the cooling mode of the samples, in temperature range of T = (10–325) K: (a) [(Na0.90K0.05Cd0.05)0.95Tb0.05]NbO3; (b) [(Na0.90K0.05Cd0.05)0.95Dy0.05]NbO3; (c) [(Na0.90K0.05Cd0.05)0.95Ho0.05]NbO3. The insets in column (1) show the dispersion and blurring of ε′/ε0 at the instant of the phase transition. The arrow shows the direction of f increase. Frequencies are denoted as follows: Ceramics 06 00021 i00896 Hz; Ceramics 06 00021 i001 8∙102 Hz; Ceramics 06 00021 i002 1.0 104 Hz; Ceramics 06 00021 i003 1.0∙105 Hz; Ceramics 06 00021 i004 2.5∙105 Hz; Ceramics 06 00021 i005 7.2∙106 Hz; Ceramics 06 00021 i006 7.2∙106 Hz; Ceramics 06 00021 i007 1.2∙107 Hz.
Figure 6. Column (1): dependence ε′/ε0, column (2): dependence ε″/ε0 on the temperature at different frequencies of the alternating electric field obtained in the cooling mode of the samples, in temperature range of T = (10–325) K: (a) [(Na0.90K0.05Cd0.05)0.95Tb0.05]NbO3; (b) [(Na0.90K0.05Cd0.05)0.95Dy0.05]NbO3; (c) [(Na0.90K0.05Cd0.05)0.95Ho0.05]NbO3. The insets in column (1) show the dispersion and blurring of ε′/ε0 at the instant of the phase transition. The arrow shows the direction of f increase. Frequencies are denoted as follows: Ceramics 06 00021 i00896 Hz; Ceramics 06 00021 i001 8∙102 Hz; Ceramics 06 00021 i002 1.0 104 Hz; Ceramics 06 00021 i003 1.0∙105 Hz; Ceramics 06 00021 i004 2.5∙105 Hz; Ceramics 06 00021 i005 7.2∙106 Hz; Ceramics 06 00021 i006 7.2∙106 Hz; Ceramics 06 00021 i007 1.2∙107 Hz.
Ceramics 06 00021 g006
Figure 7. The dependence of log(1/ε − 1/εm) on log(TTm) for the investigated ceramic SSs. The solid line shows the result of the linear approximation.
Figure 7. The dependence of log(1/ε − 1/εm) on log(TTm) for the investigated ceramic SSs. The solid line shows the result of the linear approximation.
Ceramics 06 00021 g007
Figure 8. The dependence of ε′/ε0 (upon cooling), α, and ∆L/L on the temperature of ceramic SSs: (a) Na0.90K0.05Cd0.05NbO3; (b) [(Na0.90K0.05Cd0.05)0.95La0.05]NbO3.
Figure 8. The dependence of ε′/ε0 (upon cooling), α, and ∆L/L on the temperature of ceramic SSs: (a) Na0.90K0.05Cd0.05NbO3; (b) [(Na0.90K0.05Cd0.05)0.95La0.05]NbO3.
Ceramics 06 00021 g008
Figure 9. The dependence of ε′/ε0 (upon cooling), α, and ∆L/L on the temperature of ceramic SSs: (a) [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3; (b) [(Na0.90K0.05Cd0.05)0.95Tb0.05]NbO3.
Figure 9. The dependence of ε′/ε0 (upon cooling), α, and ∆L/L on the temperature of ceramic SSs: (a) [(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3; (b) [(Na0.90K0.05Cd0.05)0.95Tb0.05]NbO3.
Ceramics 06 00021 g009
Figure 10. The dependence of ε′/ε0 (upon cooling), α, and ∆L/L on the temperature of ceramic SSs: (a) [(Na0.90K0.05Cd0.05)0.95Dy0.05]NbO3; (b) [(Na0.90K0.05Cd0.05)0.95Ho0.05]NbO3.
Figure 10. The dependence of ε′/ε0 (upon cooling), α, and ∆L/L on the temperature of ceramic SSs: (a) [(Na0.90K0.05Cd0.05)0.95Dy0.05]NbO3; (b) [(Na0.90K0.05Cd0.05)0.95Ho0.05]NbO3.
Ceramics 06 00021 g010
Figure 11. The samples of the translucent ceramics [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3, obtained by UHP. Words mean “Hello, World!” (a) the pure sample, (b) the 5 mol% sample.
Figure 11. The samples of the translucent ceramics [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3, obtained by UHP. Words mean “Hello, World!” (a) the pure sample, (b) the 5 mol% sample.
Ceramics 06 00021 g011
Figure 12. The PL spectra of the initial Na0.90K0.05Cd0.05NbO3 ceramics (1) and the modified ceramics: (2)—Ho, (3)—Tb, (4)—Dy, (5)—La, (6)—Pr. T = 300 K; λex = 325 nm.
Figure 12. The PL spectra of the initial Na0.90K0.05Cd0.05NbO3 ceramics (1) and the modified ceramics: (2)—Ho, (3)—Tb, (4)—Dy, (5)—La, (6)—Pr. T = 300 K; λex = 325 nm.
Ceramics 06 00021 g012
Table 1. The symmetry and parameters of the perovskite cell of the initial and modified by the REE Na0.9K0.05Cd0.05NbO3 SS; experimental (ρexp.) and relative (ρrel.) densities of the ceramic samples.
Table 1. The symmetry and parameters of the perovskite cell of the initial and modified by the REE Na0.9K0.05Cd0.05NbO3 SS; experimental (ρexp.) and relative (ρrel.) densities of the ceramic samples.
CompositionSymmetrya = c, Åb, Åβ (°)V, Å3ρexp. g/cm3ρrel., %
Na0.9K0.05Cd0.05NbO3M3.9223.89390.2659.94.5999.6
[(Na1-x-yKxCdy)0.95La0.05]NbO3PsC3.921 60.34.58100
[(Na1-x-yKxCdy)0.95Pr0.05]NbO3PsC3.916 60.034.62100
[(Na1-x-yKxCdy)0.95Tb0.05]NbO3M3.9213.90390.4360.04.5498.2
[(Na1-x-yKxCdy)0.95Dy0.05]NbO3M3.9213.89690.5459.94.77102
[(Na1-x-yKxCdy)0.95Ho0.05]NbO3M3.9213.89990.5459.964.73102
Table 2. The mechanical characteristics of the ceramic Na0.90K0.05Cd0.05NbO3 SS and the [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3 SSs.
Table 2. The mechanical characteristics of the ceramic Na0.90K0.05Cd0.05NbO3 SS and the [(Na0.90K0.05Cd0.05)0.95REE0.05]NbO3 SSs.
CompositionYoung’s Modulus, E, GPaMicrohardness, H, GPaCrack Resistance, Kic, MPa m0.5
Na0.90K0.05Cd0.05NbO3297.8 ± 5.44.15 ± 0.492.34 ± 0.15
[(Na0.90K0.05Cd0.05)0.95La0.05]NbO3302.0 ± 3.26.11 ± 0.622.06 ± 0.2
[(Na0.90K0.05Cd0.05)0.95Pr0.05]NbO3347.3 ± 4.38.42 ± 1.332.2 ± 0.7
[(Na0.90K0.05Cd0.05)0.95Tb0.05]NbO3352.3 ± 5.94.47 ± 0.72.32 ± 0.35
[(Na0.90K0.05Cd0.05)0.95Dy0.05]NbO3369.9 ± 4.25.3 ± 0.962.02 ± 0.1
[(Na0.90K0.05Cd0.05)0.95Ho0.05]NbO3320.8 ± 5.54.85 ± 0.591.86 ± 0.2
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

Konstantin, A.; Palatnikov, M.; Shilkina, L.; Nagaenko, A.; Shcherbina, O.; Smirnov, M.; Sidorov, N.; Kubrin, S.; Rudskaya, A.; Rudskiy, D.; et al. Multifunctional Materials Based on the Solid Solutions of the (Na, K, Cd0.5)NbO3 (KNN-Cd) System Modified with Rare Earth Elements: Crystal Structure, Microstructure, and Macroresponses. Ceramics 2023, 6, 342-363. https://doi.org/10.3390/ceramics6010021

AMA Style

Konstantin A, Palatnikov M, Shilkina L, Nagaenko A, Shcherbina O, Smirnov M, Sidorov N, Kubrin S, Rudskaya A, Rudskiy D, et al. Multifunctional Materials Based on the Solid Solutions of the (Na, K, Cd0.5)NbO3 (KNN-Cd) System Modified with Rare Earth Elements: Crystal Structure, Microstructure, and Macroresponses. Ceramics. 2023; 6(1):342-363. https://doi.org/10.3390/ceramics6010021

Chicago/Turabian Style

Konstantin, Andryushin, Mikhail Palatnikov, Lidiya Shilkina, Alexandr Nagaenko, Olga Shcherbina, Maxim Smirnov, Nikolay Sidorov, Stanislav Kubrin, Anzhela Rudskaya, Daniil Rudskiy, and et al. 2023. "Multifunctional Materials Based on the Solid Solutions of the (Na, K, Cd0.5)NbO3 (KNN-Cd) System Modified with Rare Earth Elements: Crystal Structure, Microstructure, and Macroresponses" Ceramics 6, no. 1: 342-363. https://doi.org/10.3390/ceramics6010021

Article Metrics

Back to TopTop