Next Article in Journal
Screening, Growing, and Validation by Catalog: Using Synthetic Intermediates from Natural Product Libraries to Discover Fragments for an Aspartic Protease Through Crystallography
Previous Article in Journal
Electronic Properties of Atomic Layer Deposited HfO2 Thin Films on InGaAs Compared to HfO2/GaAs Semiconductors
Previous Article in Special Issue
Investigation of the Positive Temperature Coefficient Resistivity of Nb-Doped Ba0.55Sr0.45TiO3 Ceramics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crystal Structure and Microwave Dielectric Characteristics of Novel Ba(Eu1/5Sm1/5Nd1/5Pr1/5La1/5)2Ti4O12 High-Entropy Ceramic

1
School of Electronic and Information Engineering, Hubei University of Science and Technology, Xianning 437100, China
2
Key Laboratory of Photoelectric Sensing and Intelligent Control, Hubei University of Science and Technology, Xianning 437100, China
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(9), 754; https://doi.org/10.3390/cryst14090754
Submission received: 3 August 2024 / Revised: 22 August 2024 / Accepted: 23 August 2024 / Published: 25 August 2024
(This article belongs to the Special Issue Crystal Structure and Dielectric Properties of Ceramics)

Abstract

:
High-permittivity Ba(Eu1/5Sm1/5Nd1/5Pr1/5La1/5)2Ti4O12 (BESNPLT) high-entropy ceramics (HECs) were synthesized via a solid-state route. The microstructure, sintering behavior, phase structure, vibration modes, and microwave dielectric characteristics of the BESNPLT HECs were thoroughly investigated. The phase structure of the BESNPLT HECs was confirmed to be a single-phase orthorhombic tungsten-bronze-type structure of Pnma space group. Permittivity (εr) was primarily influenced by polarizability and relative density. The quality factor (Q×f) exhibited a significant correlation with packing fraction, whereas the temperature coefficient (TCF) of the BESNPLT HECs closely depended on the tolerance factor and bond valence of B-site. The BESNPLT HECs sintered at 1400 °C, demonstrating high relative density (>97%) and optimum microwave dielectric characteristics with TCF = +38.9 ppm/°C, Q×f = 8069 GHz (@6.1 GHz), and εr = 87.26. This study indicates that high-entropy strategy was an efficient route in modifying the dielectric characteristics of tungsten-bronze-type microwave ceramics.

1. Introduction

As vital materials for microwave transmission lines, oscillators, dielectric antennas, dielectric resonators, and filters utilized in wireless fidelity (WIFI), global positioning system (GPS), Bluetooth, wireless local area network (WLAN), and 5G/6G communication systems, microwave dielectric ceramics (MWDCs) have attracted sustained attention from researchers [1,2,3,4,5]. In these applications, the microwave dielectric characteristics of MWDC must be considered, including permittivity (εr), quality factor (Q×f, Q = 1/tanδ), and temperature coefficient (TCF or τf) [6,7,8]. In particular, a suitable εr is required, as well as a near-zero TCF and a high Q×f [9,10,11]. However, the simultaneous achievement of high εr, near-zero TCF, and high Q×f value represents a significant challenge in the field of MWDC.
Recently, the idea of high entropy was applied to functional ceramics, giving rise to the emergence of high-entropy ceramics (HECs) [12,13]. These new materials are single phase, comprising a minimum of five anions and/or cations occupying a single site within the crystal [14,15]. Extensive research has been conducted on HECs in order to enhance their properties [16]. HECs with low thermal conductivity [17], excellent mechanical properties [18], dielectric properties [19], piezoelectric properties [20], ferroelectric properties [21], and corrosion resistance [22] could be fabricated by precisely adjusting the composition within a wide range. These HECs have promising potential for various applications, including piezoelectricity [23], thermal barrier coatings [24], and energy storage [25]. Considering that dielectric properties of MWDC are highly sensitive to even the slightest alterations in structure and composition, high-entropy strategy might offer a novel methodology for tuning the dielectric characteristics. Xiang et al. [26] initially documented the finding of a high-entropy Li(Lu0.2Yb0.2Er0.2Ho0.2Gd0.2)GeO4 MWDC with an orthorhombic olivine structure. Subsequently, numerous other high-entropy MWDC exhibiting optimal dielectric characteristics have been investigated [27,28,29,30,31,32,33,34,35]. Chen et al. [27] designed and synthesized (Hf1/4Zr1/4Sn1/4Ti1/4)O2 HECs, and Q×f increased significantly from 20,026 to 74,600 GHz by introduction of Hf4+ and Sn4+ into ZrTiO4. According to Lin et al. [28], the TCF of SrLaAlO4 MWDC could be adjusted from −32 to −6 ppm/°C via the introduction of equimolar Eu, Gd, Nd, and Sm in the La-site. Lai et al. [29] successfully synthesized non-equimolar (Mg1/5Co1/5Ni1/5Li2/5Zn1/5)Al2O4 HECs with TCF = −64 ppm/°C, Q×f = 58,200 GHz, and εr = 7.4. The above research mainly focuses on medium permittivity and ow-permittivity high-entropy MWDC, while there are scarce reports regarding high-entropy, high-permittivity, tungsten-bronze-type MWDC.
Previous investigations have demonstrated that BaLa2Ti4O12 MWDC with a tungsten-bronze-type structure exhibits high permittivity (εr = 95.6) for miniaturized device applications. However, low-Q×f value (2102 GHz) and high-positive TCF (+352 ppm/°C) restrict its practical application [36]. Inspired by the concept of high entropy, the equimolar Eu3+, Sm3+, Nd3+, and Pr3+ cations were incorporated into the A1-site (La-site) of BaLa2Ti4O12 MWDC, forming HECs (with the nominal composition Ba(Eu1/5Sm1/5Nd1/5Pr1/5La1/5)2Ti4 O12 to improve the Q×f and temperature stability. In this work, Ba(Eu1/5Sm1/5Nd1/5Pr1/5 La1/5)2Ti4O12 (abbreviated as BESNPLT) HECs were fabricated through a solid-state route, and the microstructure, sintering behavior, phase structure, vibration modes, and microwave dielectric characteristics of the BESNPLT HECs were thoroughly investigated. The polarizability per molar volume (αthe/Vm), packing fraction (P.F.), tolerance factor (t), and B-site bond valence (VB) were analyzed to determine the relationship between the microwave dielectric characteristics of the BESNPLT HECs.

2. Materials and Methods

Ba(Eu1/5Sm1/5Nd1/5Pr1/5La1/5)2Ti4O12 HECs were synthesized through a solid-state route utilizing Eu2O3 (99.99%), Nd2O3 (99.99%), Pr6O11 (99.9%), Sm2O3 (99.99%), La2O3 (99.99%), BaCO3 (99.8%), and TiO2 (99.84%). To remove moisture from the powders, Pr6O11 was calcined for three hours at 600 °C, while Eu2O3, Nd2O3, La2O3, and Sm2O3 were calcined for the same period at 900 °C. The initial oxide and carbonate were weighed in accordance with the stoichiometry of Ba(Eu1/5Sm1/5Nd1/5Pr1/5La1/5)2Ti4O12 (BESNPLT) and then were wet-milled in nylon jars (with deionized water and ZrO2 balls) using a speed of 280 rpm for 2 h. The mixed powders were dried at 120 °C/8 h and calcined at 1175 °C/3 h to synthesize BESNPLT powder. Following calcination, the BESNPLT powder was subjected to a second wet-milling process and subsequently dried at 120 °C/8 h. The BESNPLT powder was combined with PVA (10 wt %) and pressed into cylindrical billets and flake billets. Finally, the BLPNSET billets (cylindrical and flake) were sintered at 1350–1500 °C/4 h.
The bulk density of the BESNPLT HECs was measured by the Archimedean drainage method. The phase structure of the BESNPLT ceramic was determined utilizing a BRUKER X-ray diffraction (XRD, USA, D8 advance), which has a detection range of 10°–90°. The evaluation of the natural, sintered surface of the BESNPLT ceramic specimens was conducted employing a ZEISS Scanning Electron Microscope (SEM, GEMINI300). The element distribution of the ceramic specimens was implemented via the utilization of an energy dispersive spectrometer (EDS, Oxford, X-MAX 50, UK). Raman spectra (50~850 cm−1) of the BESNPLT HECs were obtained using a Alpha300R Raman scattering spectrometer (WITec, λ = 532 nm). Microwave dielectric characteristics of the BESNPLT HECs were evaluated at 5–6 GHz utilizing a P9375A network analyzer (USA, 0.3 MHz–26.5 GHz, Keysight) through the application of the Hakki–Coleman method [37]. The temperature coefficient (TCF) was determined using Equation (1):
TCF = ( f 90 f 30 ) × 10 6 ( 90 30 ) f 30 ( ppm / ° C )
where f90 and f30 represent the resonant frequency at 90 °C and 30 °C, respectively.

3. Results and Discussion

The XRD patterns of the BESNPLT HECs sintered at varying temperatures (1350–1500 °C) are recorded in Figure 1. BESNPLT HECs consist of Sm3+ (1.079 Å), Eu3+ (1.066 Å), Pr3+ (1.126 Å), Nd3+ (1.109 Å), and La3+ (1.16 Å) in equal molar ratios, and its average ionic radius of lanthanide elements (1.108 Å) is close to that of Nd3+ (1.109 Å) [38]. It is, therefore, anticipated that the positions of the diffraction peaks will align closely with those of BaNd2Ti4O12, as illustrated in Figure 1. The observed diffraction peaks of the BESNPLT specimens align with BaNd2Ti4O12 (PDF#44–0061), and no additional peaks are detected, indicating that the BESNPLT HECs form an orthorhombic tungsten-bronze-type phase with Pnma space group.
To gain a comprehensive insight into the structural characteristics of the BESNPLT HECs, Rietveld refinements were conducted utilizing GSAS application [39,40], with the XRD data (2θ = 10–90°) obtained from BESNPLT specimens that had been sintered at 1400 °C. The homotopic Ba4.5Nd9Ti18O54 structure (ICSD # 95269) is employed in an initial model [41], and corresponding refined results for BESNPLT HECs are illustrated in Figure 2a and Table 1. There was satisfactory concordance between the fitted (black solid line) and observed patterns (red dot) with low values of Rp = 7.29%, Rwp = 9.18%, and χ2 = 2.015 (Figure 2a and Table 1). The calculated cell volume (V) of the BESNPLT HECs sintered at 1400 °C is 2100.788 Å3 (a = 22.3607 Å, b = 7.6992 Å, c = 12.2024 Å). Figure 2b presents a structural diagram of the BESNPLT HECs projected along [001]. As depicted in Figure 2b, tungsten-bronze-type BESNPLT HECs comprises a three-dimensional framework of a vertex-sharing Ti-O octahedron interconnected at the vertices to yield three distinct types of cavities: small triangular cavities (C), diamond cavities (A1), and pentagonal cavities (A2) [42]. The small triangular cavities (C-site) are empty. The Ba2+ cations occupy the large pentagonal cavities (A2-site), whereas the remaining Ba2+ ions are distributed among the diamond cavities (A1-site), sharing these with the lanthanides (Sm3+, Eu3+, Pr3+, Nd3+, and La3+).
Raman spectroscopy represents an effective methodology for the reflection of phase structural characteristics, cation distribution, and the identification of the defects of MWDC through the detection of bond vibration characteristics [43]. The theoretically Raman-active modes of Ba4.5Nd9Ti18O54 with a Pnma space group are 258 ( Γ Raman = 73 A g + 56 B 1 g + 73 B 2 g + 56 B 3 g ). BESNPLT HECs with an identical space group (Pnma) also exhibit analogous Raman-active modes. However, the actual measurement of Raman-active modes was found to be less than the theoretical value, which was influenced by a number of factors, including peak overlap, background, and low intensity [44,45]. The Raman spectra (50 cm−1 ~ 850 cm−1) of the BESNPLT HECs at varying temperatures (1350–1500 °C) are illustrated in Figure 3, and 13 Raman-active modes (around 80, 115, 133, 181, 231, 279, 298, 330, 404, 432, 530, 592, and 751 cm−1, respectively) can be identified. As illustrated in Figure 3, the Raman-active modes remained largely unaltered with increased temperature, demonstrating that the structure of the BESNPLT HECs remained stable within the sintering temperature (S.T.) range. The Raman-active modes (80, 115, 133, and 181 cm−1) within 50–200 cm−1 are associated with the A1 and A2-site cation (Ba2+, Sm3+, Eu3+, Pr3+, Nd3+, and La3+) translation of the BESNPLT HECs [46]. The modes (231, 279, 298, and 330 cm−1) within the 200–400 cm−1 are derived from the rotation and tilt of titanium–oxygen (Ti-O) octahedron [47]. The Raman-active modes at 404 and 432 cm−1 are ascribed to bending the vibration of the titanium–oxygen bond [48]. The modes at 530 (Ag) and 592 cm−1 (B1g) have been identified as stretching the vibration of the titanium–oxygen bond. The Raman-active mode at 751 cm−1 shows a correlation with second-order scattering [49]. Furthermore, Raman spectroscopic analysis has confirmed the formation of tungsten-bronze-type BESNPLT HECs.
Figure 4 exhibits the relative density (ρre) and bulk density (ρbu) of the BESNPLT HECs at varying temperatures (1350–1500 °C). The ρre of the BESNPLT HECs could be evaluated through Equation (2):
ρ r e = ρ b u ρ t h × 100 %
where ρre represents the theoretical density, with the value calculated using XRD refinement data presented in Table 1. As the S.T. increases, the ρbu of the BESNPLT HECs increases, achieving a maximum of 5.603 g/cm3 at 1400 °C (97.34% of the theoretical value of 5.756 g/cm3). Thereafter, the ρbu decreases as the S.T. exceeds 1400 °C.
SEM photographs of the as-fired BESNPLT HECs at varying temperatures (1350–1500 °C) are represented in Figure 5a–d. The grains of the BESNPLT HECs were observed to exhibit a rod-like morphology that was in accordance with the known characteristics of tungsten-bronze-type Ba6-3xR8+2xTi18O54 (BRT, R = Sm, Nd, La, and Pr) ceramics [46,47,48,49]. As the S.T. rises, the average grain size of the BESNPLT HECs increases markedly as a consequence of grain growth. The formation of the uniform and dense (ρre > 95%) surface morphology of the BESNPLT HECs (see Figure 5a–c) was observed at 1350–1450 °C. Nevertheless, a further increase in the S.T. caused the formation of abnormal grains, characterized by the presence of pores (see Figure 5d). As illustrated in Figure 6, the EDS mapping demonstrates a highly uniform distribution of Sm3+, Eu3+, Pr3+, Nd3+, and La3+, indicating that the lanthanide ions have succeeded in entering the A1-site in the crystal lattice. According to the EDS analysis results, the barium, titanium, oxygen, and lanthanide elements were consistent with the stoichiometry of the BESNPLT HECs, and the contents of the lanthanide elements (Eu3+, Sm3+, Nd3+, Pr3+, and La3+) could be regarded as approaching equimolar proportions, with a ratio of 1.97:2.06:2.1:2.03:2.03. Consistent with the XRD and Raman analysis results, the EDS analysis also confirms the formation of the BESNPLT HEC’s solid solution.
The measured εr and ρre of the BESNPLT HECs at varying temperatures (1350–1500 °C) are illustrated in Figure 7a. As the S.T. increased, the measured εr of the BESNPLT HECs exhibited an initial increase from 85.96 to 87.26, followed by a subsequent decrease to 82.68. The tendency for the measured εr of the BESNPLT HECs to align with that of ρre indicates that density played a pivotal role in determining permittivity. The polarizability per molar volume (αthe/Vm) and porosity-corrected permittivity (εcor) of the BESNPLT HECs are presented in Figure 7b. The εcor values were determined using Equation (3) [50]:
ε c o r = ε r ( 1.5 p + 1 )
where p represents porosity determined from ρre (p = 1 − ρre). The αthe/Vm values were calculated using the Clausius–Mosotti formula, as outlined in Equation (4) [51]:
α t h e V m = 3 ( ε r 1 ) 4 π ( ε r + 2 )
where Vm represents molar volume (Vm = V/Z, Z = 2 in BESNPLT HECs). The theoretical polarizability (αthe) of the BESNPLT HECs was determined using Equation (5) [51]:
αthe(BESNPLT) = 4.5α(Ba2+) + 1.8α(La3+) + 1.8α(Pr3+) + 1.8(Nd3+) + 1.8α(Sm3+) + 1.8α(Eu3+) + 54α(O2−) + 18α(Ti4+)
where α(Eu3+) = 4.53 Å 3 , α(Sm3+) = 4.74 Å 3 , α(Pr3+) = 5.32 Å 3 , α(La3+) = 6.07 Å 3 , α(Nd3+) = 5.01 Å 3 , α(Ba2+) = 6.4 Å 3 , α(O2−) = 2.01 Å 3 , and α(Ti4+) = 2.93 Å 3 .
Furthermore, a direct proportional relationship was observed between εcor and αthe/Vm when the latter was considered as a single variable. As displayed in Figure 7b, αthe/Vm increases from 0.22493 (1350 °C) to 0.22495 (1400 °C) and then decreases to 0.2247 (1500 °C), explaining the change in the εcor. The results demonstrate that the key factors influencing the permittivity of the BESNPLT HECs are the αthe/Vm and density.
Figure 8a illustrates the Q×f and ρre of the BESNPLT HECs sintered at varying temperatures (1350–1500 °C). As presented in Figure 8a, Q×f demonstrates a slight increase from 7812 GHz to 8069 GHz as the S.T. rises from 1350 °C to 1400 °C. However, when the S.T. is further increased to 1500 °C, a notable reduction in the Q×f is observed. Dielectric loss (tanδ = 1/Q) measured at the microwave band of ceramic materials is primarily influenced by intrinsic factors, including the lattice vibration, atomic packing fraction, and lattice energy, along with extrinsic factors such as microstructure, density, and lattice defects. A comparable trend was identified in the variation of Q×f in comparison to ρre (as illustrated in Figure 8a), implying that density had a significant impact on the Q×f values of the BESNPLT HECs. In addition, the influence of the packing fraction (P.F.) on Q×f should be considered. The P.F. of the BESNPLT HECs could be determined through Equation (6) [52]:
P . F . ( % ) = Packed   ions   volume Unit   cell   volume × Z = 4 π 3 [ 4.5 r B a 3 + 1.8 ( r E u 3 + r S m 3 + r N d 3 + r Pr 3 + r L a 3 ) + 18 r T i 3 + 54 r O 3 ] V × 2
where r B a (1.52 Å), r E u (1.066 Å), r S m (1.079 Å), r N d (1.109 Å), r P r (1.126 Å), r L a (1.16 Å), r O (1.4 Å), and r T i (0.605 Å) are ionic radii of the BESNPLT HECs. The increase in the P.F. of the microwave dielectric ceramics was found to be correlated with a reduction in lattice and non-harmonic vibration, which resulted in an enhancement of Q×f. As illustrated in Figure 8b, the variation of Q×f for the BESNPLT HECs was aligned with that of the P.F. Therefore, the density and P.F are the key factors in determining the Q×f of the BESNPLT HECs.
In general, the TCF of the tungsten-bronze-type BRT (R represents the lanthanides) ceramics was found to be influenced by a number of factors, including the composition, the second phase, bond valence, and the tilt of Ti–O octahedron [53,54,55]. Figure 9a illustrates the TCF and B-site (Ti-site) bond valence (VB) of the BESNPLT HECs as a function of the S.T. As the S.T. increased, the TCF of the BESNPLT HECs demonstrated a slight decrease from 39.7 to 38.9 ppm/°C, which was subsequently followed by an increase to 47.9 ppm/°C. The VB of the BESNPLT HECs at varying temperatures (1350–1500 °C) is evaluated through Equations (7) and (8) [28]:
V j k = k v j k
v j k = exp [ R j k d j k 0.37 ]
where Rjk represents bond valence parameter, and djk denotes bond length. The decreasing degree of tilting on the Ti–O octahedron resulted in an increase in the TCF. Additionally, the degree of tilting observed in the Ti–O octahedron was influenced by the length and strength of the bonds between the oxygen and Ti-site cation. Accordingly, the degree of tilting on the Ti–O octahedron can be determined by the bond valence of the Ti-site cation, and the VB of the BESNPLT HECs at varying temperatures demonstrated an inverse variation in the TCF, as illustrated in Figure 9a.
The tolerance factor (t) was found to be significantly correlated with the degree of tilting on the Ti–O octahedron in the tungsten-bronze-type BRT (R represents the lanthanides) MWDC. The t of the BESNPLT HECs could be determined using Equation (9):
t = [ 0.18   ( R E u 3 + + R S m 3 + + R N d 3 + + R P r 3 + + R L a 3 + ) + 0.1   R B a 2 + ] + R O 2 2 ( R T i 4 + + R O 2 )
where R B a 2 + (1.52 Å), R E u 3 + (1.066 Å), R S m 3 + (1.079 Å), R N d 3 + (1.109 Å), R P r 3 + (1.126 Å), R L a 3 + (1.16 Å), R O 2 (1.4 Å), and R T i 4 + (0.605 Å) are the ionic radii of the BESNPLT HECs. Figure 9b presents the TCF and t of the BaLa2Ti4O12 (BLT) [36], BaPr2Ti4O12 (BPT) [56], BaNd2Ti4O12 (BNT) [57], BaSm2Ti4O12 (BST) [58], and BESNPLT ceramics. As t decreased, the degree of tilting on the Ti–O octahedron increased, resulting in a reduction in the TCF of the BaR2Ti4O12 (R = lanthanides) ceramics. The TCF of the BESNPLT HECs could be adjusted to an appropriate value (+38.9 ppm/°C) through high-entropy design.

4. Conclusions

In this study, high-permittivity BESNPLT HECs were successfully produced using the concept of high entropy by a solid-phase reaction route. The results of the XRD and Raman analysis demonstrated that the BESNPLT HECs belonged to a single-phase tungsten-bronze-type structure with a Pnma space group within the S.T. range of 1350–1500 °C. The EDS mapping demonstrated that the distributions of the five lanthanide elements (Eu3+, Sm3+, Nd3+, Pr3+, and La3+) were uniform in the BESNPLT HECs, confirming the formation of the BESNPLT solid solution. The microwave dielectric characteristics of the BESNPLT HECs were significantly influenced by a number of factors, including the polarizability, relative density, packing fraction, tolerance factor, and bond valence of the B-site (Ti-site). Specifically, the BESNPLT HECs achieved superior dielectric performances of TCF = +38.9 ppm/°C, Q×f = 8069 GHz (@6.1 GHz), and εr = 87.26. The present study has established a paradigm to optimize the dielectric characteristics of tungsten-bronze-type MWDC utilizing a high-entropy strategy.

Author Contributions

Conceptualization, Q.W. and G.W.; investigation, H.W. and Z.L.; methodology, Q.W.; supervision, Z.L.; validation, G.X.; writing—original draft, G.W.; writing—review and editing, H.W. and G.X. All authors have read and agreed to the published version of the manuscript.

Funding

The present work was supported by the Hubei Province Natural Science Foundation of China (2022CFB369), Hubei Province Department of Education Foundation (D20212803), and Ph.D. foundation (Grant No. BK202109).

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shehbaz, M.; Du, C.; Zhou, D.; Xia, S.; Xu, Z. Recent progress in dielectric resonator antenna: Materials, designs, fabrications, and their performance. Appl. Phys. Rev. 2023, 10, 021303. [Google Scholar] [CrossRef]
  2. Du, K.; Zhou, M.; Li, C.; Yin, C.; Cai, Y.; Cheng, M.; Zhu, W.; Wei, G.; Wang, S.; Lei, W. Ultralow-Permittivity and Temperature-Stable Ba1−xCaxMg2Al6Si9O30 Dielectric Ceramics for C-Band Patch Antenna Applications. ACS Appl. Mater. Interfaces 2024, 16, 23505–23516. [Google Scholar] [CrossRef]
  3. Wang, W.; Wang, X.; Bao, J.; Jiang, J.; Fang, Z.; Jin, B.; Shi, Z.; Darwish, M.A.; Chen, Y.; Liang, Q.; et al. Low-permittivity BaCuSi4O10-based dielectric Ceramics: An available solution to connect low temperature cofired ceramic technology and millimeter-wave communications. Chem. Eng. J. 2024, 494, 153172. [Google Scholar] [CrossRef]
  4. Sebastian, M.; Ubic, R.; Jantunen, H. Low-loss dielectric ceramic materials and their properties. Int. Mater. Rev. 2015, 60, 392–412. [Google Scholar] [CrossRef]
  5. Xiang, H.; Zhang, Y.; Chen, J.; Zhou, Y.; Tang, Y.; Chen, J.; Fang, L. Structure evolution and τf influence mechanism of Bi1−xHoxVO4 microwave dielectric ceramics for LTCC applications. J. Mater. Sci. Technol. 2024, 197, 1–8. [Google Scholar] [CrossRef]
  6. Wang, G.; Fu, Q.; Zha, L.; Hu, M.; Huang, J.; Zheng, Z.; Luo, W. Microwave dielectric characteristics of tungsten bronze-type Ba4Nd28/3Ti18-yGa4y/3O54 ceramics with temperature stable and ultra-low loss. J. Eur. Ceram. Soc. 2022, 42, 154–161. [Google Scholar] [CrossRef]
  7. Tian, H.; Zhang, X.; Zhang, Z.; Liu, Y.; Wu, H. Low-permittivity LiLn(PO3)4 (Ln = La, Sm, Eu) dielectric ceramics for microwave/millimeter-wave communication. J. Adv. Ceram. 2024, 13, 602–620. [Google Scholar] [CrossRef]
  8. Wang, G.; Zhang, D.; Huang, X.; Rao, Y.; Yang, Y.; Gan, G.; Lai, Y.; Xu, F.; Li, J.; Liao, Y. Crystal structure and enhanced microwave dielectric properties of Ta5+ substituted Li3Mg2NbO6 ceramics. J. Am. Ceram. Soc. 2020, 103, 214–223. [Google Scholar]
  9. Guo, W.; Lu, Y.; Ma, Z.; Wu, H.; Yue, Z. Defect-related broadband dielectric loss mechanisms of Na1/2Sm1/2Ti1−z(Al1/2Nb1/2)zO3 ceramics. Acta Mater. 2023, 255, 119093. [Google Scholar] [CrossRef]
  10. Yao, G.; Zhao, J.; Lu, Y.; Liu, H.; Pei, C.; Ding, Q.; Chen, M.; Zhang, Y.; Li, D.; Wang, F. Microwave dielectric properties of Li3TiO3F oxyfluorides ceramics. Crystals 2023, 13, 897. [Google Scholar] [CrossRef]
  11. Ma, L.; Tian, G.; Xiao, H.; Jiang, L.; Du, Q.; Li, H.; Yang, B. A novel spinel-type Mg3Ga2SnO8 microwave dielectric ceramic with low ε and low loss. J. Eur. Ceram. Soc. 2024, 44, 5731–5737. [Google Scholar] [CrossRef]
  12. Oses, C.; Toher, C.; Curtarolo, S. High-entropy ceramics. Nat. Rev. Mater. 2020, 5, 295–309. [Google Scholar] [CrossRef]
  13. Xiang, H.; Xing, Y.; Dai, F.-z.; Wang, H.; Su, L.; Miao, L.; Zhang, G.; Wang, Y.; Qi, X.; Yao, L. High-entropy ceramics: Present status, challenges, and a look forward. J. Adv. Ceram. 2021, 10, 385–441. [Google Scholar] [CrossRef]
  14. Akrami, S.; Edalati, P.; Fuji, M.; Edalati, K. High-entropy ceramics: Review of principles, production and applications. Mater. Sci. Eng. R Rep. 2021, 146, 100644. [Google Scholar] [CrossRef]
  15. Anandkumar, M.; Trofimov, E. Synthesis, properties, and applications of high-entropy oxide ceramics: Current progress and future perspectives. J. Alloy. Compd. 2023, 960, 170690. [Google Scholar] [CrossRef]
  16. Jiao, Y.; Dai, J.; Fan, Z.; Cheng, J.; Zheng, G.; Grema, L.; Zhong, J.; Li, H.-F.; Wang, D. Overview of high-entropy oxide ceramics. Mater. Today 2024, 77, 92–117. [Google Scholar] [CrossRef]
  17. Banerjee, R.; Chatterjee, S.; Ranjan, M.; Bhattacharya, T.; Mukherjee, S.; Jana, S.S.; Dwivedi, A.; Maiti, T. High-entropy perovskites: An emergent class of oxide thermoelectrics with ultralow thermal conductivity. ACS Sustain. Chem. Eng. 2020, 8, 17022–17032. [Google Scholar] [CrossRef]
  18. Harrington, T.J.; Gild, J.; Sarker, P.; Toher, C.; Rost, C.M.; Dippo, O.F.; McElfresh, C.; Kaufmann, K.; Marin, E.; Borowski, L. Phase stability and mechanical properties of novel high entropy transition metal carbides. Acta Mater. 2019, 166, 271–280. [Google Scholar] [CrossRef]
  19. Zhou, S.; Pu, Y.; Zhang, Q.; Shi, R.; Guo, X.; Wang, W.; Ji, J.; Wei, T.; Ouyang, T. Microstructure and dielectric properties of high entropy Ba(Zr0.2Ti0.2Sn0.2Hf0.2Me0.2)O3 perovskite oxides. Ceram. Int. 2020, 46, 7430–7437. [Google Scholar] [CrossRef]
  20. Li, H.; Zhao, J.; Li, Y.; Chen, L.; Chen, X.; Qin, H.; Zhou, H.; Li, P.; Guo, J.; Wang, D. Bismuth Ferrite-Based Lead-Free High-Entropy Piezoelectric Ceramics. ACS Appl. Mater. Interfaces 2024, 16, 9078–9087. [Google Scholar] [CrossRef]
  21. Zhang, M.; Xu, X.; Ahmed, S.; Yue, Y.; Palma, M.; Svec, P.; Gao, F.; Abrahams, I.; Reece, M.J.; Yan, H. Phase transformations in an Aurivillius layer structured ferroelectric designed using the high entropy concept. Acta Mater. 2022, 229, 117815. [Google Scholar] [CrossRef]
  22. von Fieandt, K.; Paschalidou, E.-M.; Srinath, A.; Soucek, P.; Riekehr, L.; Nyholm, L.; Lewin, E. Multi-component (Al, Cr, Nb, Y, Zr) N thin films by reactive magnetron sputter deposition for increased hardness and corrosion resistance. Thin Solid Film 2020, 693, 137685. [Google Scholar] [CrossRef]
  23. Liu, Y.; Yang, J.; Deng, S.; Zhang, Y.; Zhang, Y.; Sun, S.; Wang, L.; Jiang, X.; Huo, C.; Liu, H. Flexible polarization configuration in high-entropy piezoelectrics with high performance. Acta Mater. 2022, 236, 118115. [Google Scholar] [CrossRef]
  24. Ren, K.; Wang, Q.; Shao, G.; Zhao, X.; Wang, Y. Multicomponent high-entropy zirconates with comprehensive properties for advanced thermal barrier coating. Scr. Mater. 2020, 178, 382–386. [Google Scholar] [CrossRef]
  25. Guo, J.; Yu, H.; Ren, Y.; Qi, H.; Yang, X.; Deng, Y.; Zhang, S.-T.; Chen, J. Multi-symmetry high-entropy relaxor ferroelectric with giant capacitive energy storage. Nano Energy 2023, 112, 108458. [Google Scholar] [CrossRef]
  26. Xiang, H.; Yao, L.; Chen, J.; Yang, A.; Yang, H.; Fang, L. Microwave dielectric high-entropy ceramic Li(Gd0.2Ho0.2Er0.2Yb0.2Lu0.2)GeO4 with stable temperature coefficient for low-temperature cofired ceramic technologies. J. Mater. Sci. Technol. 2021, 93, 28–32. [Google Scholar] [CrossRef]
  27. Ding, Y.H.; Liu, L.; Guo, R.Z.; Li, L.; Chen, X.M. (Hf0.25Zr0.25Sn0.25Ti0.25)O2 high-entropy ceramics and their microwave dielectric characteristics. J. Am. Ceram. Soc. 2022, 105, 6710–6717. [Google Scholar] [CrossRef]
  28. Lin, F.L.; Liu, B.; Hu, C.C.; Song, K.X. Novel high-entropy microwave dielectric ceramics Sr(La0.2Nd0.2Sm0.2Eu0.2Gd0.2)AlO4 with excellent temperature stability and mechanical properties. J. Eur. Ceram. Soc. 2023, 43, 2506–2512. [Google Scholar] [CrossRef]
  29. Xie, M.; Li, X.; Lai, Y.; Qi, C.; Yin, J.; Gong, W.; Li, Y.; Liu, Q.; Wu, C. Phase evolution and microware dielectric properties of high-entropy spinel-type (Mg0.2Co0.2Ni0.2Li0.4Zn0.2)Al2O4 ceramics. J. Eur. Ceram. Soc. 2024, 44, 284–292. [Google Scholar] [CrossRef]
  30. Chen, D.; Yan, N.; Cao, X.; Li, F.; Liu, L.; Shen, Q.; Zhou, H.; Li, C. Entropy regulation in LaNbO4-based fergusonite to implement high-temperature phase transition and promising dielectric properties. J. Adv. Ceram. 2023, 12, 1067–1080. [Google Scholar] [CrossRef]
  31. Huanrong, T.; Zhang, X.; Du, W.; Feng, Z.; Wang, L.; Haitao, W.; Xia, W. Structure characteristics and microwave/terahertz dielectric response of low-permittivity (La0.2Nd0.2Sm0.2Eu0.2Gd0.2)2Zr3(Mo O4)9 high-entropy ceramics. Ceram. Int. 2024, 50, 6403–6411. [Google Scholar]
  32. Xiao, L.; Deng, L.; Li, Y.; Qing, Z.; Xi, Y.; Zhu, J.; Peng, S. A TiSnNbTaGa2O12 high-entropy microwave dielectric ceramic with rutile structure and near-zero τf. J. Eur. Ceram. Soc. 2024, 44, 277–283. [Google Scholar] [CrossRef]
  33. Chen, D.; Zhu, X.; Xiong, S.; Zhu, G.; Liu, L.; Khaliq, J.; Li, C. Tunable Microwave Dielectric Properties in Rare-Earth Niobates via a High-Entropy Configuration Strategy to Induce Ferroelastic Phase Transition. ACS Appl. Mater. Interfaces 2023, 15, 52776–52787. [Google Scholar] [CrossRef]
  34. Liu, K.; Zhang, H.; Liu, C.; Li, J.; Shi, L.; Wang, X.; Zhang, D. Crystal structure and microwave dielectric properties of (Mg0.2Ni0.2Zn0.2Co0.2Mn0.2)2SiO4-A novel high-entropy ceramic. Ceram. Int. 2022, 48, 23307–23313. [Google Scholar] [CrossRef]
  35. Du, Q.; Duan, J.; Ma, L.; Jiang, L.; Li, H. Sintering temperature, microwave dielectric properties and structural characteristics of Li2Zn6MgTi6O20 ceramic with high configuration entropy. J. Eur. Ceram. Soc. 2024, 44, 5639–5645. [Google Scholar] [CrossRef]
  36. Muhammad, R.; Iqbal, Y.; Rambo, C.R. Characterization of Ba4.5Re9Ti18O54 (Re = La, Nd) microwave dielectric ceramics. J. Mater. Sci. Mater. Electron. 2014, 25, 1652–1656. [Google Scholar] [CrossRef]
  37. Hakki, B.; Coleman, P.D. A dielectric resonator method of measuring inductive capacities in the millimeter range. IRE Trans. Microw. Theory Tech. 1960, 8, 402–410. [Google Scholar] [CrossRef]
  38. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A Cryst. Phys. Diffr. Theor. Gen. Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  39. Rietveld, H.M. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
  40. Toby, B.H. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef]
  41. Tang, C.; Roberts, M.; Azough, F.; Leach, C.; Freer, R. Synchrotron X-ray diffraction study of Ba4.5Nd9Ti18O54 microwave dielectric ceramics at 10–295 K. J. Mater. Res. 2002, 17, 675–682. [Google Scholar] [CrossRef]
  42. Ohsato, H. Science of tungstenbronze-type like Ba6−3xR8+2xTi18O54 (R = rare earth) microwave dielectric solid solutions. J. Eur. Ceram. Soc. 2001, 21, 2703–2711. [Google Scholar] [CrossRef]
  43. Hao, S.Z.; Zhou, D.; Pang, L.X. The spectra analysis and microwave dielectric properties of [Ca0.55(Sm1−xBix)0.3]MoO4 ceramics. J. Am. Ceram. Soc. 2019, 102, 3103–3109. [Google Scholar] [CrossRef]
  44. Li, L.; Wang, X.; Luo, W.; Wang, S.; Yang, T.; Zhou, J. Internal-strain-controlled tungsten bronze structural ceramics for 5G millimeter-wave metamaterials. J. Mater. Chem. C 2021, 9, 14359–14370. [Google Scholar] [CrossRef]
  45. Ning, F.; Lin, G.; Yuan, S.; Qi, Z.; Jiang, J.; Zhang, T. Correlation between vibrational modes of A-site ions and microwave dielectric properties in (1 − x)CaTiO3−x(Li0.5Sm0.5)TiO3 ceramics. J. Alloys Comp. 2017, 729, 742–748. [Google Scholar] [CrossRef]
  46. Guo, W.; Ma, Z.; Luo, Y.; Chen, Y.; Yue, Z.; Li, L. Structure, defects, and microwave dielectric properties of Al-doped and Al/Nd co-doped Ba4Nd9.33Ti18O54 ceramics. J. Adv. Ceram. 2022, 11, 629–640. [Google Scholar] [CrossRef]
  47. Wang, G.; Fu, Q.; Guo, P.; Hu, M.; Wang, H.; Yu, S.; Zheng, Z.; Luo, W. Crystal structure, spectra analysis and dielectric characteristics of Ba4M28/3Ti18O54 (M = La, Pr, Nd, and Sm) microwave ceramics. Ceram. Int. 2021, 47, 1750–1757. [Google Scholar] [CrossRef]
  48. Luo, W.; Wang, X.; Bai, B.; Qiao, J.; Chen, X.; Wen, Y.; Sun, J.; Li, L.; Zhou, J. B-site internal-strain regulation engineering of tungsten bronze structural dielectric ceramics. J. Am. Ceram. Soc. 2024, 107, 367–376. [Google Scholar] [CrossRef]
  49. Guo, W.; Yang, Y.; Ma, Z.; Lu, Y.; Yue, Z. Effects of Zr4+ and Hf4+ substitution on the structure and dielectric response of Ba4Sm9. 33Ti18O54 ceramics. Ceram. Int. 2024, 50, 12443–12449. [Google Scholar] [CrossRef]
  50. Penn, S.J.; Alford, N.M.; Templeton, A.; Wang, X.; Xu, M.; Reece, M.; Schrapel, K. Effect of porosity and grain size on the microwave dielectric properties of sintered alumina. J. Am. Ceram. Soc. 1997, 80, 1885–1888. [Google Scholar] [CrossRef]
  51. Shannon, R.D. Dielectric polarizabilities of ions in oxides and fluorides. J. Appl. Phys. 1993, 73, 348–366. [Google Scholar] [CrossRef]
  52. Kim, E.S.; Chun, B.S.; Freer, R.; Cernik, R.J. Effects of packing fraction and bond valence on microwave dielectric properties of A2+B6+O4 (A2+: Ca, Pb, Ba; B6+: Mo, W) ceramics. J. Eur. Ceram. Soc. 2010, 30, 1731–1736. [Google Scholar] [CrossRef]
  53. Ubic, R.; Reaney, I.; Lee, W. Microwave dielectric solid–solution phase in system BaO-Ln2O3-TiO2 (Ln = lanthanide cation). Int. Mater. Rev. 1998, 43, 205–219. [Google Scholar] [CrossRef]
  54. Xiong, Z.; Tang, B.; Fang, Z.; Yang, C.; Zhang, S. Crystal structure, Raman spectroscopy and microwave dielectric properties of Ba3.75Nd9.5Ti18−z(Al1/2Nb1/2)zO54 ceramics. J. Alloys Compd. 2017, 723, 580–588. [Google Scholar] [CrossRef]
  55. Wang, G.; Fu, Q.; Shi, H.; Tian, F.; Guo, P.; Yan, L.; Yu, S.; Zheng, Z.; Luo, W. Novel thermally stable, high quality factor Ba4(Pr0.4Sm0.6)28/3Ti18−yGa4y/3O54 microwave dielectric ceramics. J. Am. Ceram. Soc. 2020, 103, 2520–2527. [Google Scholar] [CrossRef]
  56. Solomon, S.; Santha, N.; Jawahar, I.; Sreemoolanadhan, H.; Sebastian, M.; Mohanan, P. Tailoring the microwave dielectric properties of BaRE2Ti4O12 and BaRE2Ti5O14 ceramics by compositional variations. J. Mater. Sci. Mater. Electron. 2000, 11, 595–602. [Google Scholar] [CrossRef]
  57. Ubic, R.; Reaney, I.M.; Lee, W.E.; Samuels, J.; Evangelinos, E. Properties of the microwave dielectric phase Ba6−3XNd8+ 2XTi18O54. Ferroelectrics 1999, 228, 271–282. [Google Scholar] [CrossRef]
  58. Cheng, P.-S.; Yang, C.-F.; Chen, Y.-C.; Tzou, W.-C. The reaction sequence and dielectric properties of BaSm2Ti4O12 ceramics. Ceram. Int. 2000, 26, 877–881. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of BESNPLT HECs sintered at 1350–1500 °C.
Figure 1. XRD patterns of BESNPLT HECs sintered at 1350–1500 °C.
Crystals 14 00754 g001
Figure 2. (a) Rietveld refinement plots of BESNPLT HECs. (b) Structural diagram of BESNPLT HECs.
Figure 2. (a) Rietveld refinement plots of BESNPLT HECs. (b) Structural diagram of BESNPLT HECs.
Crystals 14 00754 g002
Figure 3. Raman spectra of BESNPLT HECs sintered at varying temperatures.
Figure 3. Raman spectra of BESNPLT HECs sintered at varying temperatures.
Crystals 14 00754 g003
Figure 4. ρre and ρbu of BESNPLT HECs sintered at 1350–1500 °C.
Figure 4. ρre and ρbu of BESNPLT HECs sintered at 1350–1500 °C.
Crystals 14 00754 g004
Figure 5. SEM morphology of BESNPLT HECs sintered at (a) 1350 °C; (b) 1400 °C; (c) 1450 °C; (d) 1500 °C.
Figure 5. SEM morphology of BESNPLT HECs sintered at (a) 1350 °C; (b) 1400 °C; (c) 1450 °C; (d) 1500 °C.
Crystals 14 00754 g005
Figure 6. EDS mapping of BESNPLT HECs.
Figure 6. EDS mapping of BESNPLT HECs.
Crystals 14 00754 g006
Figure 7. (a) εr and ρre of BESNPLT HECs; (b) εcor and αthe/Vm of BESNPLT HECs.
Figure 7. (a) εr and ρre of BESNPLT HECs; (b) εcor and αthe/Vm of BESNPLT HECs.
Crystals 14 00754 g007
Figure 8. (a) Q×f and ρre of BESNPLT HECs; (b) Q×f and packing fraction of BESNPLT HECs.
Figure 8. (a) Q×f and ρre of BESNPLT HECs; (b) Q×f and packing fraction of BESNPLT HECs.
Crystals 14 00754 g008
Figure 9. (a) TCF and B-site bond valence of BESNPLT HECs; (b) BaLa2Ti4O12 (BLT) [36], BaPr2Ti4O12 (BPT) [56], BaNd2Ti4O12 (BNT) [57], BaSm2Ti4O12 (BST) [58], and BESNPLT ceramics.
Figure 9. (a) TCF and B-site bond valence of BESNPLT HECs; (b) BaLa2Ti4O12 (BLT) [36], BaPr2Ti4O12 (BPT) [56], BaNd2Ti4O12 (BNT) [57], BaSm2Ti4O12 (BST) [58], and BESNPLT ceramics.
Crystals 14 00754 g009
Table 1. The refinement parameters of BESNPLT HECs.
Table 1. The refinement parameters of BESNPLT HECs.
BLPNSET1350 °C1400 °C1450 °C1500 °C
a (Å)22.355722.360722.366722.3747
b (Å)7.70117.69927.70097.7005
c (Å)12.203112.202412.204812.2061
V3)2100.9342100.7882102.2172103.079
α = β = γ90909090
Rp (%)6.657.297.336.92
Rwp (%)8.489.189.448.96
χ2 (%)1.7392.0152.052.19
ρtheo (g/cm3)5.75595.75635.75245.7501
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

Wan, Q.; Li, Z.; Wang, H.; Xiong, G.; Wang, G. Crystal Structure and Microwave Dielectric Characteristics of Novel Ba(Eu1/5Sm1/5Nd1/5Pr1/5La1/5)2Ti4O12 High-Entropy Ceramic. Crystals 2024, 14, 754. https://doi.org/10.3390/cryst14090754

AMA Style

Wan Q, Li Z, Wang H, Xiong G, Wang G. Crystal Structure and Microwave Dielectric Characteristics of Novel Ba(Eu1/5Sm1/5Nd1/5Pr1/5La1/5)2Ti4O12 High-Entropy Ceramic. Crystals. 2024; 14(9):754. https://doi.org/10.3390/cryst14090754

Chicago/Turabian Style

Wan, Qing, Zeping Li, Huifeng Wang, Gang Xiong, and Geng Wang. 2024. "Crystal Structure and Microwave Dielectric Characteristics of Novel Ba(Eu1/5Sm1/5Nd1/5Pr1/5La1/5)2Ti4O12 High-Entropy Ceramic" Crystals 14, no. 9: 754. https://doi.org/10.3390/cryst14090754

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