Next Article in Journal
Enhancement of Resistive and Synaptic Characteristics in Tantalum Oxide-Based RRAM by Nitrogen Doping
Previous Article in Journal
Non-Noble Metal Catalysts in Cathodic Oxygen Reduction Reaction of Proton Exchange Membrane Fuel Cells: Recent Advances
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cu2+-Ion-Substitution-Driven Microstructure and Microwave Dielectric Properties of Mg1−xCuxAl2O4 Ceramics

1
School of Mechanical and Electrical Engineering, Chengdu University of Technology, Chengdu 610059, China
2
Guangdong Provincial Key Laboratory of Electronic Functional Materials and Devices, Huizhou University, Huizhou 516001, China
3
State Key Laboratory of Electronic Thin Films and Integrated Devices, University of Electronic Science and Technology of China, Chengdu 610054, China
4
Science and Technology on Combustion and Explosion Laboratory, Xi’an Modern Chemistry Research Institute, Xi’an 710065, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(19), 3332; https://doi.org/10.3390/nano12193332
Submission received: 17 August 2022 / Revised: 13 September 2022 / Accepted: 21 September 2022 / Published: 24 September 2022

Abstract

:
In this work, Cu-substituted MgAl2O4 ceramics were prepared via solid-state reaction. The crystal structure, cation distribution, and microwave dielectric properties of Mg1−xCuxAl2O4 ceramics were investigated. Cu2+ entered the MgAl2O4 lattice and formed a spinel structure. The substitution of Cu2+ ions for Mg2+ ions contributed to Al3+ ions preferential occupation of the octahedron and changed the degree of inversion. The quality factor (Qf) value, which is correlated with the degree of inversion, increased to a maximum value at x = 0.04 and then decreased. Ionic polarizability and relative density affected the dielectric constant (εr) value. The temperature coefficient of the resonant frequency (τf) value, which was dominated by the total bond energy, generally shifted to the positive direction. Satisfactory microwave dielectric properties were achieved in x = 0.04 and sintered at 1550 °C: εr = 8.28, Qf = 72,800 GHz, and τf = −59 ppm/°C. The Mg1−xCuxAl2O4 solid solution, possessing good performance, has potential for application in the field of modern telecommunication technology.

1. Introduction

The microwave dielectric ceramics have been extensively applied in various fields, including fifth-generation wireless systems, intelligent transmission systems, and ultrahigh-speed wireless local area networks [1,2,3]. In the application of millimeter waves, there is an urgent requirement for microwave dielectric ceramics with excellent performance in the following areas: a high quality factor (Qf) to enable microwave frequency selectivity, a low dielectric constant (εr) to shorten the delay time of signal propagation, and a near-zero temperature coefficient of resonant frequency (τf) to ensure the stability of frequency against temperature changes [4,5]. However, few single-phase materials can meet these requirements simultaneously due to the mutual restrictions of the three parameters. In general, ideal εr and Qf values can be obtained by selecting material systems, whereas the near-zero τf value is tailored through two materials with opposite τf values [6,7,8,9]. However, this approach tends to deteriorate the εr and Qf values.
Previous studies have found that the τf value is related to octahedral distortion in some ceramic crystals [10,11]. Microwave dielectric ceramics with superior τf values can be obtained by adjusting the distortion of the octahedron without deteriorating the εr and Qf values [4]. Therefore, it is important to improve the microwave dielectric properties utilizing the crystal structure. In general, the spinel structural degrees of freedom, such as the cell parameters, the oxygen fractional coordinates, and degrees of inversion, can be tailored via substitution [12]. MgAl2O4 is known to have a typical cubic spinel belonging to symmetry group Fd-3m (227); the molecular formula is [Mg1−λAlλ]IV[Al2−λMgλ]VIO4, where λ value, which is related to the degree of inversion [13,14], represents the occupation of Al3+ cations at tetrahedral site.
MgAl2O4 ceramic, which generally exhibits a Qf value of ~68,900 GHz (the highest Qf value is over 200,000 [15]) and a low εr value (~8.75), is one of the candidate material for a millimeter-wave communication substrate [16]. However, it has a large negative τf value (~−75 ppm/°C). It has been reported that MgAl2O4-based composite dielectric ceramics, such as MgAl2O4-TiO2 and MgAl2O4-(Ca0.8Sr0.2)TiO3, have near-zero τf values [6,7]. However, the εr and Qf values are also deteriorated in this system. It is worth mentioning that the τf value in MgAl2O4 can be improved through the crystal structure [17,18]. Previously, in MgAl2O4 ceramics, it has been shown that the enhancement in Qf value corresponds to the cation distribution [19]. Moreover, the degree of inversion in MgAl2O4 ceramics, prepared by the solid-state reaction or molten-salt reaction routes, has also been investigated [15]. A high degree of inversion represents a high Qf value, and the preferential occupation of Al3+ could enhance the covalency of M-O bonds in a [MO4] tetrahedron of MgAl2O4 (M = Mg and Al). Consequently, the cation distribution of Al3+ in MgAl2O4 can be discussed to ameliorate the microwave dielectric properties.
In general, the ionic radius of Cu2+ ion is close to that of Mg2+ ions [4,20,21], which is beneficial for forming Mg1−xCuxAl2O4 solid solutions. Additionally, a significant Jahn–Teller effect can be observed when Cu2+ ions occupy the octahedral site in a spinel structure [22]. This can contribute to the regulation of the microstructure of MgAl2O4. In addition, CuO can also reduce the sintering temperature of ceramics [4,21]. Therefore, in this work, the Cu2+ ion was considered as a substitution of the Mg2+ ion in MgAl2O4. Mg1−xCuxAl2O4 ceramics were synthesized through a solid-state route. The phase composition, microstructure, and microwave dielectric properties were investigated in detail.

2. Experimental Procedure

Mg1−xCuxAl2O4 (x = 0, 0.04, 0.08, 0.12, 0.16, and 0.20) ceramics were synthesized via the conventional solid-state route. Analytic-grade purity MgO, CuO, and Al2O3 powders (Shanghai Macklin Biochemical Co., Ltd., Shanghai, China) with the range of particle size at 45–80 μm were used as starting materials, which were weighed and wet-mixed in deionized water using zirconia balls in a plastic container at 300 rpm for 4 h. The obtained slurries were dried and calcined in alumina crucibles at 1450 °C for 4 h. Subsequently, the calcined powders were ground into a fine form and pressed under a uniaxial pressure of 10 MPa into cylindrical disks with 12 mm diameter and 5–6 mm height. Samples were sintered at temperature levels ranging from 1450 to 1600 °C for 4 h.
To confirm the crystalline phase of the Mg1−xCuxAl2O4 ceramics, X-ray diffraction (XRD, Miniflex600, Rigaku, Tokyo, Japan), using Cu Kα radiation (λ = 1.54 Å) at room temperature, was measured in the 2θ angle range between 10° and 120° with a step of 0.01°, and counting time for 5 s per step. Based on the XRD results, the crystal structure was analyzed using the Rietveld refinement method using FullProf software (FullProf Suite May2021 64b, The FullProf Team, Grenoble, France) [23]. The microstructures and morphologies of the samples were analyzed by a scanning electron microscope (SEM, JSM-6490; JEOL, Tokyo, Japan) at an accelerating voltage of 20 kV. The Al3+ ion distributions of Mg1−xCuxAl2O4 were investigated through 27Al solid-state magic-angle spinning nuclear magnetic resonance (MAS-NMR) with a spinning frequency of 12 kHz (Avance II 600 MHz, Bruker, Fällanden, Switzerland).
Microwave dielectric properties (εr, Qf, and τf) of these samples were measured by the Hakki–Coleman dielectric resonator with a vector network analyzer (N5230A, Agilent Technologies, Santa Clara, CA, USA). The τf value was calculated based on the resonant frequencies at 25 and 85 °C:
τ f = f 85 f 25 60 × f 25 × 10 6 ( ppm / ° C )
where f t represents the resonant frequency at t °C.

3. Results and Discussion

Figure 1 shows the microwave dielectric properties of Mg1−xCuxAl2O4 ceramics sintered at 1450–1600 °C. Good microwave dielectric properties were obtained at x = 0.04 with sintering at 1550 °C: εr = 8.28, Qf = 72,800 GHz, and τf = −59 ppm/°C. With the sintering temperature at 1550 °C, the τf value experienced a significant increase in the negative direction to about −59 ppm/°C at 0 ≤ x ≤ 0.04; then, the rapid fall was witnessed and a steady rise was observed at 0.08 ≤ x ≤ 0.20. Figure 1b shows the εr value, which increased first up to x = 0.12 and then presented a modest downward trend when the sintering temperatures were 1450 and 1500 °C, whereas it remained virtually unchanged at 1550 and 1600 °C. In addition, it is well known that Qf values are determined by both intrinsic and extrinsic factors. Intrinsic factor is mainly caused by lattice vibration, while extrinsic factor is dominated by grain boundary, secondary phase, and densification [24,25]. With the increase in the x value, the Qf values of the samples with different sintering temperatures increased initially and then showed a steady drop. The maximum Qf value was acquired at x = 0.04 with sintering at 1550 and 1600 °C. Compared with previous studies (see Table 1) [6,15,17,18,26,27,28,29,30,31,32,33], the sintering temperature and τf value of this work can be improved. However, Qf is lower than the best reported value [15], one of the reasons may be the different experimental conditions, such as preparation method, sintering temperature and ball milling time, etc. To understand the microstructure and microwave dielectric properties of Mg1−xCuxAl2O4 ceramics, the phase composition, relative density, and cation distribution were investigated in this study.
XRD analysis is often carried out to identify phases. The XRD patterns of Mg1−xCuxAl2O4 ceramics sintered at 1550 °C are illustrated in Figure 2. All the diffraction peaks can be assigned to those for the standard MgAl2O4 (PDF # 21-1125) pattern with a space group Fd-3m (227). Meanwhile, there is no apparent peak corresponding to any additional secondary phase containing Cu or structural phase transitions observed from Figure 2, indicating the successful formation of Mg1−xCuxAl2O4 solid solutions [34].
Figure 3 displays the Rietveld refinement of the XRD patterns for Mg1−xCuxAl2O4 ceramics, and the refinement results are presented in Table 2. Both the Bragg positions models are well within the standard indexed peaks, indicating that the refinement result is acceptable. According to the refinement, the Cu2+ ions occupied the tetrahedral site, with the exception of a small amount of the octahedral site. Figure 4 shows the schematic diagram of a crystal structure from MgAl2O4 to Mg1−xCuxAl2O4. The Mg2+ and Cu2+ ions significantly occupy the 8a Wyckoff position, and the Al3+ ions mainly occupy the 16d Wyckoff position. They form a [Mg(T)/Cu(T)O4] tetrahedron and an [Al(M)O6] octahedron, respectively. In general, Al3+ ions preferentially occupy tetrahedra, which can effectively improve the Qf value (~232,301 GHz) [19,26]. However, Al3+ ions mainly occupy the octahedron, which is consistent with low Qf values (~72,800 GHz) of the Mg1−xCuxAl2O4 system. The cell parameters showed a nonlinear trend with an increase in Cu2+ ions content; that is, they first increased, then decreased, and finally increased. On the one hand, the radius of a Cu2+ ion (r = 0.73 Å) is slightly larger than that of a Mg2+ ion (r = 0.72 Å) [20], which may have led to the increase in cell parameters. On the other hand, the Cu2+ ions’ octahedral coordination has a significant Jahn–Teller effect. The Jahn–Teller distortion can enhance the polarizing effect of Cu2+ ions [22]. Therefore, the hybridization of Cu2+ ions is responsible for a decrease in average cell parameters [22]. Consequently, the two mechanisms were in competition with each other, resulting in a nonlinear variation trend of cell parameters. The cation distribution, which is a significant variation, had no obvious effect on the microwave dielectric properties.
For further analysis of the effects of the introduction of Cu2+ ions on microwave dielectric properties, the variations in the cation distribution in Al3+ were measured via the 27Al solid-state MAS-NMR measurement, which was used to evaluate the Al3+ sites in Mg1−xCuxAl2O4 ceramics. Figure 5a shows the 27Al NMR spectra of Mg1−xCuxAl2O4 ceramics sintered at 1550 °C. The spectra indicate three signals with chemical shifts at ca. 10, 17, and 71 ppm. They correspond to octahedrally coordinated aluminum (AlO6), pentahedrally coordinated aluminum (AlO5), and tetrahedrally coordinated aluminum (AlO4), respectively [35,36]. For the emergence of AlO5, a dynamic disorder occurred between the twisted tetrahedral structure and octahedral structure and froze some Al ions stuck in the pentahedral structure at high temperatures [37]. In Figure 5a, the peak intensities of AlO4 at 71 ppm gradually weakened, whereas the peaks of AlO6 at 10 ppm and AlO5 at 17 ppm first increased and then decreased with the increase in Cu2+ content. This result indicated that the redistribution of Al3+ ions in the lattice occurred by the substitution of Cu2+ for Mg2+. Moreover, the peaks transferred to low chemical shifts on account of the second-order quadrupolar-order shifts with the increase in Cu2+ content [31,38]. On the whole, the peak intensity of AlO6 is significantly larger than that of AlO4, which indicates that Al3+ ions mainly occupied octahedral sites. This was in accordance with the XRD results. The different peaks indicated the existence of an intermediate spinel structure in the system. The intermediate spinel structure can be described as [Mg1−λ2+Alλ3+]IV[Al2−λ3+Mgλ2+]VIO4, where λ is the degree of the inversion of spinel structure, corresponding to the fraction of Al3+ ions in the tetrahedral site. The value of λ ranges from 0 (normal spinel: (Mg2+)IV(Al23+)VIO4) to 1 (inverse spinel: (Al3+)IV(Al3+Mg2+)VIO4) [13,14]. It is known that the microwave dielectric properties are related to λ [14,15,39]. The value of λ can be calculated with the following formula [40]:
λ = 2 I ( AlO 4 ) I ( AlO 4 ) + I ( AlO 6 )
where I(AlO4) and I(AlO6) are the intensities of tetrahedral and octahedral resonances, respectively. The λ values are displayed in Figure 5b. Considering the trend of λ in Figure 5b, the Al3+ cations preferentially occupied the octahedral sites. It has been reported that the preferential tetrahedron site occupation of Al3+ could enhance the Qf value of the system [19,26]. Consequently, when 0.04 < x < 0.16, the Qf values reduce with the decrease in the λ value in Mg1−xCuxAl2O4 ceramics (see Figure 5b). In addition, the entry of Cu2+ ions into the MgAl2O4 lattice can lead to disordered charge distribution, which can cause a decrease in Qf value at high Cu2+ ions content [41].
The relative density can be calculated according to measured and theoretical density. The theoretical density can be derived from XRD refinements. The results are presented in Table 2. The relative density first increased and then gradually decreased, and the maximum value, which was obtained at x = 0.04, was 95.59%. The SEM shows that the densification was consistent with the relative density (see Figure S1). It also indicates that the moderate amount of CuO can promote the sintering of MgAl2O4 ceramics, which is beneficial for obtaining uniform and dense microstructures (see Figure S1). The relationship between the εr value and the relative density is presented in Figure 6, which shows that the εr value and relative density showed the same trend when x ≤ 0.12. With the increase in Cu2+ at x > 0.12, the pores in ceramics also played an important role for the εr value. To further investigate the effect of porosity on the εr value, the porosity-corrected dielectric constant (εrc) can be calculated by a spherical-pore model [42]:
ε r = ε rc ( 1 3 P ( ε rc 1 ) 2 ε rc + 1 )
where εr and P (1 − ρrelative) are the measured εr and the porosity, respectively. The calculated results are listed in Table 3. The εrc value was higher than the εr value, which indicated that the air trapped in the pores contributed to the decrease in the dielectric constant [43]. It is worth mentioning that, with the increase in Cu2+ at x > 0.12, the relative density maintained a declining trend, while the εr value had a slight growth. In response to this difference, apart from the relative density and pores, the variation in the εr value can be evaluated by the Clausius–Mosotti equation [44]:
ε theo = 3 V m + 8 π α theo 3 V m 4 π α theo
where Vm and αtheo represent the molecular volume and theoretical ionic polarizabilities, respectively. αtheo can be calculated as follows [45,46]:
α theo = ( 1 x ) α ( M g 2 + ) + x α ( C u 2 + ) + 2 α ( A l 3 + ) + 4 α ( O 2 )
where the αi value corresponds to the individual ionic dielectric polarizabilities. The results are listed in Table 3. The theoretical ionic polarizabilities of Mg1−xCuxAl2O4 ceramics, which increased linearly from 10.94 to 11.10, are shown in Figure 6. In general, the increase in polarizabilities led to the increase in the εr value, and the expected variation only occurred at x ≥ 0.12 [47]. This indicates that the ionic polarizabilities have a more significant impact on the εr value than that of the relative density at x > 0.12.
In order to clarify the correlation between the chemical bonds and the microwave dielectric properties of Mg1−xCuxAl2O4 ceramics at 1550 °C, the complex chemical bond theory analysis was carried out, which was contributed by Phillips, Van Vecten, and Levine (P-V-L) [48,49,50,51,52]. The detailed process of the P-V-L theory analysis is presented in the Supplementary Materials. The bond length, lattice energy, and bond energy, which are calculated through P-V-L theory, are shown in Tables S1–S4, respectively. The τf value is the combined result of the bonding strength and the crystal structure. In general, the binding force between the ions in the unit cell was stronger, the restoring force that affected the tilt of the oxygen octahedron was higher, the unit cell was less affected at high temperatures, and the τf value was closer to zero [21,53]. Figure 7 shows the τf value and the total bond energy as a function of the x value. When x ≤ 0.04 and x ≥ 0.08, the τf value shifted to zero with the increase in total bond energy, indicating that the system tended to be stable.

4. Conclusions

A single-phase Mg1−xCuxAl2O4 ceramic with a spinel structure was formulated and analyzed. The Cu2+ ions occupied the tetrahedral site, whereas the Al3+ ions preferentially occupied octahedral site, resulting in a low the Qf value. In addition, the entry of Cu2+ ions into the MgAl2O4 lattice lead to disordered charge distribution, which can cause a decrease in Qf value at high Cu2+ ions content. The Cu substitution had the high bond energy, which contributed to the temperature stability of the samples at x ≤ 0.04 and x ≥ 0.08. Then, the τf value moved toward the positive direction. Good microwave dielectric properties were achieved at x = 0.04, sintered at 1550 °C: εr = 8.28, Qf = 72,800 GHz, and τf = −59 ppm/°C. Therefore, the Qf and τf values of the Mg1xCuxAl2O4 solid solution were improved, maintaining a low εr value. This study suggests that Mg1xCuxAl2O4 is a promising candidate ceramic, possessing a high Qf value and a low dielectric constant, for use in modern communication systems.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/nano12193332/s1, Figure S1: SEM micrograph for Mg1−xCuxAl2O4 ceramics: (a) x = 0, 1550 °C; (b) x = 0.04, 1550 °C; (c) x = 0.12, 1550 °C; (d) x = 0.04, 1450 °C; (e) x = 0.04, 1500 °C; and (f) x = 0.04, 1600 °C. Table S1: Bond length d (Å) for Mg1−xCuxAl2O4 ceramics sintered at 1550 °C;. Table S2: Parameters of lattice energy for Mg1−xCuxAl2O4 ceramics sintered at 1550 °C;. Table S3: Lattice energy Ucal (kJ mol−1) for Mg1−xCuxAl2O4 ceramics sintered at 1550 °C;. Table S4: Bond energy E (kJ mol−1) for Mg1−xCuxAl2O4 ceramics sintered at 1550 °C;. References [21,51,54,55,56,57,58,59,60,61,62,63] were cited in the Supplementary Materials.

Author Contributions

Conceptualization, Y.L. and M.Y.; methodology, H.L.; validation, Q.Z.; investigation, F.W. and C.W.; resources, Y.L.; data curation, B.L.; writing—original draft preparation, B.L., X.Y. and F.Y.; writing—review and editing, Y.L. and M.Y.; visualization, X.Y. and F.Y.; project administration, M.Y.; funding acquisition, Y.L., M.Y. and W.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Guangdong Provincial Key Laboratory of Electronic Functional Materials and Devices, grant number EFMD2022005Z; the Sichuan Science and Technology Program, grant number 2022NSFSC0901; the Young and Middle-aged Core Teacher Development Project of Chengdu University of Technology; and the Major Development Project of Chengdu University of Technology (Yibin Campus).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, Y.; Deng, J.Y.; Li, M.J.; Sun, D.; Guo, L.X. A MIMO Dielectric Resonator Antenna with Improved Isolation for 5G Mm-Wave Applications. IEEE Antennas Wirel. Propag. Lett. 2019, 18, 747–751. [Google Scholar] [CrossRef]
  2. Choudhary, P.; Kumar, R.; Gupta, N. Dielectric Material Selection of Microstrip Patch Antenna for Wireless Communication Applications Using Ashby’s Approach. Int. J. Microw. Wirel. Technol. 2015, 7, 579–587. [Google Scholar] [CrossRef]
  3. Zhang, Y.P.; Liu, D. Antenna-on-Chip and Antenna-in-Package Solutions to Highly Integrated Millimeter-Wave Devices for Wireless Communications. IEEE Trans. Antennas Propag. 2009, 57, 2830–2841. [Google Scholar] [CrossRef]
  4. Lai, Y.; Tang, X.; Huang, X.; Zhang, H.; Liang, X.; Li, J.; Su, H. Phase Composition, Crystal Structure and Microwave Dielectric Properties of Mg2−xCuxSiO4 Ceramics. J. Eur. Ceram. Soc. 2018, 38, 1508–1516. [Google Scholar] [CrossRef]
  5. Hu, X.; Huang, X.J.; Chen, Y.H.; Li, Y.; Ling, Z.Y. Phase Evolution and Microwave Dielectric Properties of SrTiO3 Added ZnAl2O4–Zn2SiO4–SiO2 Ceramics. Ceram. Int. 2020, 46, 7050–7054. [Google Scholar] [CrossRef]
  6. Surendran, K.P.; Bijumon, P.V.; Mohanan, P.; Sebastian, M.T. (1−x)MgAl2O4xTiO2 Dielectrics for Microwave and Millimeter Wave Applications. Appl. Phys. A Mater. Sci. Process. 2005, 81, 823–826. [Google Scholar] [CrossRef]
  7. Yu, J.; Shen, C.; Qiu, T. Effect of Microwave Sintering on the Microstructure and Dielectric Properties of 0.92MgAl2O4–0.08(Ca0.8Sr0.2)TiO3 Ceramics. J. Mater. Sci. Mater. Electron. 2015, 26, 2737–2741. [Google Scholar] [CrossRef]
  8. Nong, L.; Cao, X.; Li, C.; Liu, L.; Fang, L.; Khaliq, J. Influence of Cation Order on Crystal Structure and Microwave Dielectric Properties in xLi4/3Ti5/3O4-(1−x)Mg2TiO4 (0.6 ≤ x ≤ 0.9) Spinel Solid Solutions. J. Eur. Ceram. Soc. 2021, 41, 7683–7688. [Google Scholar] [CrossRef]
  9. Li, H.; Tang, B.; Li, Y.; Qing, Z. Effects of Mg2.05SiO4.05 Addition on Phase Structure and Microwave Properties of MgTiO3-CaTiO3 Ceramic System. Mater. Lett. 2015, 145, 30–33. [Google Scholar] [CrossRef]
  10. Jia, X.; Xu, Y.; Zhao, P.; Li, J.; Li, W. Structural Dependence of Microwave Dielectric Properties in Ilmenite-Type Mg(Ti1−xNbx)O3 Solid Solutions by Rietveld Refinement and Raman Spectra. Ceram. Int. 2021, 47, 4820–4830. [Google Scholar] [CrossRef]
  11. Liao, Q.; Li, L.; Ding, X. Phase Constitution, Structure Analysis and Microwave Dielectric Properties of Zn 0.5Ti1−xZrxNbO4 Ceramics. Solid State Sci. 2012, 14, 1385–1391. [Google Scholar] [CrossRef]
  12. Bruschini, E.; Speziale, S.; Bosi, F.; Andreozzi, G.B. Fe–Mg Substitution in Aluminate Spinels: Effects on Elastic Properties Investigated by Brillouin Scattering. Phys. Chem. Miner. 2018, 45, 759–772. [Google Scholar] [CrossRef]
  13. Millard, R.L.; Peterson, R.C.; Hunter, B.K. Temperature Dependence of Cation Disorder in MgAl2O4 Spinel Using 27Al and 17O Magic-Angle Spinning NMR. Am. Mineral. 1992, 77, 44–52. [Google Scholar]
  14. Kan, A.; Okazaki, H.; Takahashi, S.; Ogawa, H. Microwave Dielectric Properties and Cation Distribution of Spinel-Structured Mg0.4Al2.4−xGaxO4 Ceramics with Cation Defect. Jpn. J. Appl. Phys. 2018, 57, 11UE03. [Google Scholar] [CrossRef]
  15. Takahashi, S.; Kan, A.; Ogawa, H. Microwave Dielectric Properties and Crystal Structures of Spinel-Structured MgAl2O4 Ceramics Synthesized by a Molten-Salt Method. J. Eur. Ceram. Soc. 2017, 37, 1001–1006. [Google Scholar] [CrossRef]
  16. Qin, T.; Zhong, C.; Shang, Y.; Cao, L.; Wang, M.; Tang, B.; Zhang, S. Effects of LiF on Crystal Structure, Cation Distributions and Microwave Dielectric Properties of MgAl2O4. J. Alloys Compd. 2021, 886, 161278. [Google Scholar] [CrossRef]
  17. Huang, C.L.; Tai, C.Y.; Huang, C.Y.; Chien, Y.H. Low-Loss Microwave Dielectrics in the Spinel-Structured (Mg1−xNix)Al2O4 Solid Solutions. J. Am. Ceram. Soc. 2010, 93, 1999–2003. [Google Scholar] [CrossRef]
  18. Tsai, W.C.; Liou, Y.H.; Liou, Y.C. Microwave Dielectric Properties of MgAl2O4-CoAl2O4 Spinel Compounds Prepared by Reaction-Sintering Process. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2012, 177, 1133–1137. [Google Scholar] [CrossRef]
  19. Takahashi, S.; Ogawa, H.; Kan, A. Electronic States and Cation Distributions of MgAl2O4 and Mg0.4Al2.4O4 Microwave Dielectric Ceramics. J. Eur. Ceram. Soc. 2018, 38, 593–598. [Google Scholar] [CrossRef]
  20. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomie Distances in Halides and Chaleogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  21. Lai, Y.; Zeng, Y.; Han, J.; Liang, X.; Zhong, X.; Liu, M.; Duo, B.; Su, H. Structure Dependence of Microwave Dielectric Properties in Zn2−xSiO4−x-xCuO Ceramics. J. Eur. Ceram. Soc. 2021, 41, 2602–2609. [Google Scholar] [CrossRef]
  22. Le Nestour, A.; Gaudon, M.; Villeneuve, G.; Andriessen, R.; Demourgues, A. Steric and Electronic Effects Relating to the Cu2+ Jahn-Teller Distortion in Zn1−xCuxAl2O4 Spinels. Inorg. Chem. 2007, 46, 2645–2658. [Google Scholar] [CrossRef]
  23. Rodríguez-Carvajal, J. Recent Advances in Magnetic Structure Determination by Neutron Powder Diffraction. Phys. B Phys. Condens. Matter 1993, 192, 55–69. [Google Scholar] [CrossRef]
  24. Kim, J.M.; Jo, H.W.; Kim, E.S. Effect of Electronegativity on Microwave Dielectric Properties of MgTi1−x(A1/3Sb2/3)xO3 (A = Mg2+, Zn2+) Ceramics. Int. J. Appl. Ceram. Technol. 2019, 16, 2053–2059. [Google Scholar] [CrossRef]
  25. Lai, Y.; Su, H.; Wang, G.; Tang, X. Low-Temperature Sintering of Microwave Ceramics with High Qf Values through LiF Addition. J. Am. Ceram. Soc. 2019, 102, 1893–1903. [Google Scholar] [CrossRef]
  26. Zheng, C.W.; Fan, X.C.; Chen, X.M. Analysis of Infrared Reflection Spectra of (Mg1−xZnx)Al2O4 Microwave Dielectric Ceramics. J. Am. Ceram. Soc. 2008, 91, 490–493. [Google Scholar] [CrossRef]
  27. Wu, S.; Xue, J.; Fan, Y. Spinel Mg(Al, Ga)2O4 Solid Solution as High-Performance Microwave Dielectric Ceramics. J. Am. Ceram. Soc. 2014, 97, 3555–3560. [Google Scholar] [CrossRef]
  28. Qin, T.; Zhong, C.; Qin, Y.; Tang, B.; Zhang, S. The Structure Evolution and Microwave Dielectric Properties of MgAl2−x(Mg0.5Ti0.5)xO4 Solid Solutions. Ceram. Int. 2020, 46, 19046–19051. [Google Scholar] [CrossRef]
  29. Takahashi, S.; Kan, A.; Ogawa, H. Effects of Cation Distribution on Microwave Dielectric Properties of Mg1−xZnxAl2O4 Ceramics. Mater. Chem. Phys. 2017, 200, 257–263. [Google Scholar] [CrossRef]
  30. Huang, C.L.; Chien, Y.H.; Tai, C.Y.; Huang, C.Y. High-Q Microwave Dielectrics in the (Mg1−xZnx)Al2O4 (x = 0–0.1) System. J. Alloys Compd. 2011, 509, 2010–2012. [Google Scholar] [CrossRef]
  31. Takahashi, S.; Kan, A.; Ogawa, H. Microwave Dielectric Properties and Cation Distributions of Zn1−3xAl2+2xO4 Ceramics with Defect Structures. J. Eur. Ceram. Soc. 2017, 37, 3059–3064. [Google Scholar] [CrossRef]
  32. Takahashi, S.; Kan, A.; Ogawa, H. Microwave Dielectric Properties and Crystal Structures of Mg0.7Al2.2O4 and Mg0.4Al2.4O4 Ceramics with Defect Structures. J. Am. Ceram. Soc. 2017, 100, 3497–3504. [Google Scholar] [CrossRef]
  33. Fu, P.; Xu, Y.; Shi, H.; Zhang, B.; Ruan, X.; Lu, W. The Effect of Annealing Process on the Optical and Microwave Dielectric Properties of Transparent MgAl2O4 Ceramics by Spark Plasma Sintering. Opt. Mater. (Amst.) 2014, 36, 1232–1237. [Google Scholar] [CrossRef]
  34. Fregola, R.A.; Bosi, F.; Skogby, H.; Hålenius, U. Cation Ordering over Short-Range and Long-Range Scales in the MgAl2O4-CuAl2O4 Series. Am. Mineral. 2012, 97, 1821–1827. [Google Scholar] [CrossRef]
  35. Khanna, A.; Saini, A.; Chen, B.; González, F.; Pesquera, C. Structural Study of Bismuth Borosilicate, Aluminoborate and Aluminoborosilicate Glasses by 11B and 27Al MAS NMR Spectroscopy and Thermal Analysis. J. Non. Cryst. Solids 2013, 373–374, 34–41. [Google Scholar] [CrossRef]
  36. MacKenzie, K.J.D.; Smith, M.E. Chapter 5: 27Al NMR. In Pergamon Materials Series; Elsevier: Amsterdam, The Netherlands, 2002; pp. 271–330. [Google Scholar]
  37. Harindranath, K.; Anusree Viswanath, K.; Vinod Chandran, C.; Bräuniger, T.; Madhu, P.K.; Ajithkumar, T.G.; Joy, P.A. Evidence for the Co-Existence of Distorted Tetrahedral and Trigonal Bipyramidal Aluminium Sites in SrAl12O19 from 27Al NMR Studies. Solid State Commun. 2010, 150, 262–266. [Google Scholar] [CrossRef]
  38. Pellerin, N.; Dodane-Thiriet, C.; Montouillout, V.; Beauvy, M.; Massiot, D. Cation Sublattice Disorder Induced by Swift Heavy Ions in MgAl2O4 and ZnAl2O4 Spinels: 27Al Solid-State NMR Study. J. Phys. Chem. B 2007, 111, 12707–12714. [Google Scholar] [CrossRef]
  39. Takahashi, S.; Imai, Y.; Kan, A.; Hotta, Y.; Ogawa, H. Microwave Dielectric Properties of Composites Consisting of MgAl2O4 Filler Synthesized by Molten-Salt Method and Isotactic Polypropylene Polymer Matrix. Jpn. J. Appl. Phys. 2015, 54, 10NE02. [Google Scholar] [CrossRef]
  40. Blaakmeer, E.S.; Rosciano, F.; Van Eck, E.R.H. Lithium Doping of MgAl2O4 and ZnAl2O4 Investigated by High-Resolution Solid State NMR. J. Phys. Chem. C 2015, 119, 7565–7577. [Google Scholar] [CrossRef]
  41. Tamura, H. Microwave Dielectric Losses Caused by Lattice Defects. J. Eur. Ceram. Soc. 2006, 26, 1775–1780. [Google Scholar] [CrossRef]
  42. Santhosh Kumar, T.; Pamu, D. Effect of V2O5 on Microwave Dielectric Properties of Non-Stoichiometric MgTiO3 Ceramics. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2015, 194, 86–93. [Google Scholar] [CrossRef]
  43. Sayyadi-Shahraki, A.; Taheri-Nassaj, E.; Hassanzadeh-Tabrizi, S.A.; Barzegar-Bafrooei, H. A New Temperature Stable Microwave Dielectric Ceramic with Low-Sintering Temperature in Li2TiO3-Li2Zn3Ti 4O12 System. J. Alloys Compd. 2014, 597, 161–166. [Google Scholar] [CrossRef]
  44. Wang, F.; Lai, Y.; Zeng, Y.; Yang, F.; Li, B.; Yang, X.; Su, H.; Han, J.; Zhong, X. Enhanced Microwave Dielectric Properties in Mg2Al4Si5O18 Through Cu2+ Substitution. Eur. J. Inorg. Chem. 2021, 2021, 2464–2470. [Google Scholar] [CrossRef]
  45. Shannon, R.D. Dielectric Polarizabilities of Ions in Oxides and Fluorides. J. Appl. Phys. 1993, 73, 348–366. [Google Scholar] [CrossRef]
  46. Zhang, X.; Tang, B.; Fang, Z.; Yang, H.; Xiong, Z.; Xue, L.; Zhang, S. Structural Evolution and Microwave Dielectric Properties of a Novel Li3Mg2−x/3Nb1−2x/3TixO6 System with a Rock Salt Structure. Inorg. Chem. Front. 2018, 5, 3113–3125. [Google Scholar] [CrossRef]
  47. Xiao, E.C.; Shi, F.; Fu, G.; Ren, Q.; Dou, G.; Lei, W.; Qi, Z.M. Effects of BaCu(B2O5) Additives on the Crystal Structures and Dielectric Properties of CaMgGeO4 ceramics for LTCC Applications. CrystEngComm 2020, 22, 4768–4777. [Google Scholar] [CrossRef]
  48. Levine, B.F. D-Electron Effects on Bond Susceptibilities and Ionicities. Phys. Rev. B 1973, 7, 2591–2600. [Google Scholar] [CrossRef]
  49. Phillips, J.C.; Van Vechten, J.A. Dielectric Classification of Crystal Structures, Ionization Potentials, and Band Structures. Phys. Rev. Lett. 1969, 22, 705–708. [Google Scholar] [CrossRef]
  50. Levine, B.F. Bond Susceptibilities and Ionicities in Complex Crystal Structures. J. Chem. Phys. 1973, 59, 1463–1486. [Google Scholar] [CrossRef]
  51. Wu, Z.; Meng, Q. Semiempirical Study on the Valences of Cu and Bond Covalency in Y1−xCaxBa2Cu3O6+y. Phys. Rev. B-Condens. Matter Mater. Phys. 1998, 58, 958–962. [Google Scholar] [CrossRef]
  52. Yang, H.; Zhang, S.; Yang, H.; Li, E. Usage of P-V-L Bond Theory in Studying the Structural/Property Regulation of Microwave Dielectric Ceramics: A Review. Inorg. Chem. Front. 2020, 7, 4711–4753. [Google Scholar] [CrossRef]
  53. Xiao, M.; Lou, J.; Zhou, Z.; Gu, Q.; Wei, Y.; Zhang, P. Crystal Structure and Microwave Dielectric Properties of Ta5+ Substituted MgZrNb2O8 Ceramics. Ceram. Int. 2017, 43, 15567–15572. [Google Scholar] [CrossRef]
  54. Xia, W.; Li, L.; Ning, P.; Liao, Q. Relationship Between Bond Ionicity, Lattice Energy, and Microwave Dielectric Properties of Zn(Ta1−xNbx)2O6 Ceramics. J. Am. Ceram. Soc. 2012, 95, 2587–2592. [Google Scholar] [CrossRef]
  55. Van Vechten, J.A. Quantum Dielectric Theory of Electronegativity in Covalent Systems. I. Electronic Dielectric Constant. Phys. Rev. (Ser. I) 1969, 182, 891–905. [Google Scholar] [CrossRef]
  56. Zhang, P.; Zhao, Y.; Wang, X. The Relationship Between Bond Ionicity, Lattice Energy, Coefficient of Thermal Expansion and Microwave Dielectric Properties of Nd(Nb1−xSbx)O4 Ceramics. Dalton Trans. 2015, 44, 10932–10938. [Google Scholar] [CrossRef]
  57. Zhang, P.; Zhao, Y.; Li, L. The Correlations Among Bond Ionicity, Lattice Energy and Microwave Dielectric Properties of (Nd1−xLax)NbO4 Ceramics. Phys. Chem. Chem. Phys. 2015, 17, 16692–16698. [Google Scholar] [CrossRef]
  58. Li, C.; Ding, S.; Zhang, Y.; Zhu, H.; Song, T. Effects of Ni2+ Substitution on the Crystal Structure, Bond Valence, and Microwave Dielectric Properties of BaAl2–2xNi2xSi2O8–x Ceramics. J. Eur. Ceram. Soc. 2020, 41, 2610–2616. [Google Scholar] [CrossRef]
  59. Sanderson, R.T. Electronegativity and Bond Energy. J. Am. Chem. Soc. 1983, 105, 2259–2261. [Google Scholar] [CrossRef]
  60. Xiao, M.; Wei, Y.; Zhang, P. The Correlations between Complex Chemical Bond Theory and Microwave Dielectric Properties of Ca2MgSi2O7 Ceramics. J. Electron. Mater. 2019, 48, 1652–1659. [Google Scholar] [CrossRef]
  61. Xiao, M.; Sun, H.; Zhou, Z.; Zhang, P. Bond Ionicity, Lattice Energy, Bond Energy, and Microwave Dielectric Properties of Ca1−xSrxWO4 Ceramics. Ceram. Int. 2018, 44, 20686–20691. [Google Scholar] [CrossRef]
  62. Sanderson, R. Multiple and Single Bond Energies in Inorganic Molecules. J. Inorg. Nucl. Chem. 1968, 30, 375–393. [Google Scholar] [CrossRef]
  63. Luo, Y.-R. Comprehensive Handbook of Chemical Bond Energies; CRC Press: Boca Raton, FL, USA, 2007. [Google Scholar] [CrossRef]
Figure 1. The microwave dielectric properties of Mg1−xCuxAl2O4 (x = 0–0.2) ceramics sintered at 1450–1600 °C: (a) Qf value, (b) εr value, and (c) τf value.
Figure 1. The microwave dielectric properties of Mg1−xCuxAl2O4 (x = 0–0.2) ceramics sintered at 1450–1600 °C: (a) Qf value, (b) εr value, and (c) τf value.
Nanomaterials 12 03332 g001
Figure 2. The XRD patterns of the Mg1−xCuxAl2O4 ceramics.
Figure 2. The XRD patterns of the Mg1−xCuxAl2O4 ceramics.
Nanomaterials 12 03332 g002
Figure 3. The XRD with Rietveld refinements: (a) x = 0, (b) x = 0.04, (c) x = 0.08, (d) x = 0.12, (e) x = 0.16, and (f) x = 0.20.
Figure 3. The XRD with Rietveld refinements: (a) x = 0, (b) x = 0.04, (c) x = 0.08, (d) x = 0.12, (e) x = 0.16, and (f) x = 0.20.
Nanomaterials 12 03332 g003
Figure 4. The schematic diagram of a crystal structure for MgAl2O4.
Figure 4. The schematic diagram of a crystal structure for MgAl2O4.
Nanomaterials 12 03332 g004
Figure 5. (a) 27Al NMR spectra of Mg1−xCuxAl2O4 ceramics sintered at 1550 °C; (b) the effects of the degree of inversion on Qf value at 1550 °C.
Figure 5. (a) 27Al NMR spectra of Mg1−xCuxAl2O4 ceramics sintered at 1550 °C; (b) the effects of the degree of inversion on Qf value at 1550 °C.
Nanomaterials 12 03332 g005
Figure 6. The variation between εr value at 1550 °C and theoretical ionic polarizabilities; the relative density.
Figure 6. The variation between εr value at 1550 °C and theoretical ionic polarizabilities; the relative density.
Nanomaterials 12 03332 g006
Figure 7. The τf value and the total bond energy of Mg1−xCuxAl2O4 ceramics sintered at 1550 °C.
Figure 7. The τf value and the total bond energy of Mg1−xCuxAl2O4 ceramics sintered at 1550 °C.
Nanomaterials 12 03332 g007
Table 1. Microwave dielectric properties and preparation condition of MgAl2O4-based and ZnAl2O4 ceramics.
Table 1. Microwave dielectric properties and preparation condition of MgAl2O4-based and ZnAl2O4 ceramics.
Sampleτf (ppm/°C)Qf (GHz)εrTs (°C)Milling Time (h)Preparation MethodRef.
Mg0.96Cu0.04Al2O4−5972,8008.2815504solid state reactionThis work
MgAl2O4N/A82,0007.91550–170024solid state reaction[26]
ZnAl2O4N/A106,0008.61550–170024solid state reaction [26]
Mg(Al0.4Ga0.6)2O4−16107,0008.871285–15356solid state reaction[27]
MgAl1.94(Mg0.5Ti0.5)0.06O4−61.3698,0009.114256solid state reaction [28]
Mg0.25Zn0.75Al2O4−60222,6008.401600 24solid state reaction[29]
0.75MgAl2O4-0.25TiO2−12105,40011.041400–146024solid state reaction[6]
(Mg0.75Ni0.25)Al2O4−53.5130,0008.211480–160024solid state reaction[17]
(Mg0.95Zn0.05)Al2O4−64~−70156,0008.11480–160012solid state reaction[30]
Zn0.4Al2.4O4−66202,4688.21500–160024molten salt method[31]
MgAl2O4−62.4201,6907.8160024molten salt method[15]
Mg0.7Al2.2O4 −60201,1117.7160024molten salt method[32]
Mg0.4Al2.4O4−60232,3017.5160024molten salt method[32]
Mg0.8Co0.2Al2O4−6049,3008.461475–150012reaction-sintering process[18]
Transparent MgAl2O4N/A52,6408.201350N/Aspark plasma sintering[33]
Note: N/A is not applicable.
Table 2. The cell parameters, density, and reliable factors were obtained based on the Rietveld refinements of XRD.
Table 2. The cell parameters, density, and reliable factors were obtained based on the Rietveld refinements of XRD.
x = 0x = 0.04x = 0.08x = 0.12x = 0.16x = 0.20
a = b = c (Å)8.08498.08648.08698.08388.08178.0828
V3)528.467528.775528.857528.259527.841528.060
Rp (%)6.356.515.414.764.353.99
Rwp (%)8.548.887.196.105.635.17
Rexp (%)4.724.604.293.943.703.56
χ23.273.732.812.402.322.10
ρm (g∙cm−3)3.3793.4543.4653.4913.4843.503
ρt (g∙cm−3)3.5773.6143.6523.6963.7383.776
ρr (%)94.4995.5994.9094.4993.2292.79
Note: The ρm, ρt, and ρr are the measured density, the measured density, and the relative density, respectively.
Table 3. Measured dielectric constant (εr) and porosity-corrected dielectric constant (εrc) of Mg1−xCuxAl2O4 ceramics sintered at 1550 °C.
Table 3. Measured dielectric constant (εr) and porosity-corrected dielectric constant (εrc) of Mg1−xCuxAl2O4 ceramics sintered at 1550 °C.
x Valuex = 0x = 0.04x = 0.08x = 0.12x = 0.16x = 0.2
P0.0550.0440.0510.0550.0680.072
εr8.148.288.238.228.268.29
εrc8.758.768.808.839.049.12
εtheo7.797.928.058.238.408.54
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lai, Y.; Yin, M.; Li, B.; Yang, X.; Gong, W.; Yang, F.; Zhang, Q.; Wang, F.; Wu, C.; Li, H. Cu2+-Ion-Substitution-Driven Microstructure and Microwave Dielectric Properties of Mg1−xCuxAl2O4 Ceramics. Nanomaterials 2022, 12, 3332. https://doi.org/10.3390/nano12193332

AMA Style

Lai Y, Yin M, Li B, Yang X, Gong W, Yang F, Zhang Q, Wang F, Wu C, Li H. Cu2+-Ion-Substitution-Driven Microstructure and Microwave Dielectric Properties of Mg1−xCuxAl2O4 Ceramics. Nanomaterials. 2022; 12(19):3332. https://doi.org/10.3390/nano12193332

Chicago/Turabian Style

Lai, Yuanming, Ming Yin, Baoyang Li, Xizhi Yang, Weiping Gong, Fan Yang, Qin Zhang, Fanshuo Wang, Chongsheng Wu, and Haijian Li. 2022. "Cu2+-Ion-Substitution-Driven Microstructure and Microwave Dielectric Properties of Mg1−xCuxAl2O4 Ceramics" Nanomaterials 12, no. 19: 3332. https://doi.org/10.3390/nano12193332

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