Next Article in Journal
Motion Parallax Holograms Generated from an Existing Hologram
Next Article in Special Issue
Mini-Review: Recent Technologies of Electrode and System in the Enzymatic Biofuel Cell (EBFC)
Previous Article in Journal
Evaluation of the Ultimate Strength of the Ultra-High-Performance Fiber-Reinforced Concrete Beams
Previous Article in Special Issue
Identification of Transcription Factors, Biological Pathways, and Diseases as Mediated by N6-methyladenosine Using Tensor Decomposition-Based Unsupervised Feature Extraction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microwave Performance, Microstructure, and Crystallization of (Mg0.6Zn0.4)1−yNiyTiO3 Ilmenite Ceramics

1
Department of Electronic Engineering, National Yunlin University of Science and Technology, Yunlin 64002, Taiwan
2
Department of Electrical Engineering, I-Shou University, Kaohsiung City 84001, Taiwan
*
Author to whom correspondence should be addressed.
Submission received: 20 February 2021 / Revised: 20 March 2021 / Accepted: 22 March 2021 / Published: 25 March 2021
(This article belongs to the Special Issue Selected Papers from ISET 2020 and ISPE 2020)

Abstract

:
The sintering behavior, microstructure analysis, crystallization, and microwave performance of (Mg0.6Zn0.4)1−yNiyTiO3 (y = 0.01–0.2) ceramics, processed with raw powders of MgO, NiO, ZnO, and TiO2 via the conventional solid-state method, are investigated. The main phases of (Mg0.6Zn0.4)1−yNiyTiO3 ceramics were obtained. With partial replacement by Ni2+, the (Mg0.6Zn0.4)0.95Ni0.05TiO3 could be well sintered at 1200 °C, and the microwave performance was shown to be positively correlated with sintering temperature. The permittivity (εr) saturated at 18.7–19.3, and the quality factor (Qf) values approached 72,000–165,000 (GHz) as the sintering temperatures increase from 1125 to 1250 °C. The temperature coefficient of resonance frequency (τf) falls in a stable range of −62.9 to −66 ppm/°C as sintering temperature rising. A permittivity (εr) of 19.3, a maximum Qf value of 165,000 (GHz), and a temperature coefficient (τf) of −65.4 ppm/°C were measured for the samples at 1200 °C/4 h. (Mg0.6Zn0.4)0.95Ni0.05TiO3 material system shows high potential for applications of high frequency-selection components in satellite communication and 5G wireless telecommunication systems.

1. Introduction

Recent demands for high-frequency electronic components are increasing, and ceramic-based compositions are extensively used in satellite communications, including DBS TV, GPS, Internet of Things (IoT), etc., and 5G telecommunications, cell phone, cell site, etc. Furthermore, microwave dielectric ceramics are widely used for filters, resonators, and mixers in wireless communication systems [1,2,3,4]. In microwave telecommunications, microwave resonators fabricated by dielectric ceramics result in the miniaturization of microwave devices.
Materials possessing the following dielectric properties are suitable for dielectric resonators [5,6,7,8]: (1) Permittivity (εr) > 20: The permittivity is an essential factor for component size reduction. (2) Quality factor (tanδ, the inverse of the dielectric loss) > 30,000 (GHz): Lower insertion loss and steeper cut-off can be obtained due to a higher Qf value, which ensured the lower dielectric loss of components. (3) Temperature coefficient of resonance frequency (τf) < ±10 ppm/°C: A near-zero τf value (<±10 ppm/°C) is necessary for the stability of device performance even temperature changes enormously in the environment. Generally speaking, three parameters strongly affect the size, frequency selectivity, and temperature stability of the device to the system, respectively.
MgTiO3-based ceramics, which possessed an ilmenite-type structure and trigonal space group R-3 have caught much attention and typical applications as dielectric resonators [9], filters, antennas, and so on for microwave communication systems due to its good microwave dielectric performance recently. MgTiO3 sintered at 1350 °C was demonstrated excellent dielectric performance in microwave frequency: εr ~17, Qf value ~160,000 (at 7 GHz) and temperature coefficients of resonance frequency (τf) ~ −51 ppm/°C [10]. Furthermore, (Mg0.6Zn0.4)TiO3 sintered at 1200 °C with εr ~19.8, Qf value ~144,000 and τf ~ −66 ppm/°C have also been demonstrated as promising titanate candidate for broad applications in this microwave frequency. Some researchers pay attention to comprehensively investigate the effects of relative contents between Mg and Zn in (MgyZn1−y)TiO3 system [11,12]. As a result, microwave dielectric performance of (MgyZn1−y)TiO3 highly depended on the content of Mg.
It was reported that the partial replacement of Mg by B2+ (B2+ = Co, Ni, and Zn) increases the structural densification and upgrades the dielectric performance of MgTiO3. (Mg0.95B2+0.05)TiO3 with an ilmenite-type structure has been documented to possess excellent dielectric performance [8,13,14]. Since the ionic radii of B2+ (0.0745 nm of Co2+, 0.068 nm of Ni2+, and 0.082 nm of Zn2+) approach to that of Mg2+ (0.072 nm), Mg2+ ions can be replaced by B2+ ions to form (Mg, B2+)TiO3 compound. However, the (Mg0.95 B2+0.05)TiO3 shows a negative temperature coefficient (τf). In the previous study, combining some large positive τf compensator with (Mg0.95 B2+0.05)TiO3 is effective to obtain near-zero τf value in the whole material system [14,15,16,17,18,19,20,21]. It is noted that the second phase of Mg0.95B2+0.05Ti2O5 (B2+ = Co, Ni, and Zn) might deteriorate the dielectric performance of the (Mg0.95 B2+0.05)TiO3 system, and we could not precisely evaluate the effect of Mg0.95B2+0.05Ti2O5 phase [22,23,24,25,26,27].
In this paper, with partial substitution of Mg2+ (0.072 nm) by Ni2+(0.069 nm), the microstructure densification of (Mg0.6Zn0.4)0.95Ni0.05TiO3 was observed, and microwave dielectric performance of that was also further improved compared to (Mg0.6Zn0.4)TiO3. The (Mg0.6Zn0.4)0.95Ni0.05TiO3 were fabricated via a solid-state method. The microwave dielectric performance was investigated according to the densification, the x-ray diffraction (XRD) analysis, and the samples’ microstructures (SEM and EDS). The correlation between the microstructure and the Qf value was also investigated in detail.

2. Experimental Analysis

Conventional solid-state methods were utilized to synthesize samples of (Mg0.6Zn0.4)1−yNiyTiO3 (y = 0.01–0.2) from high-purity oxide powders (>99.9%): MgO, NiO, and TiO2. Raw MgO powder was fired at 600 °C/1h to avoid moisture retention due to its hygroscopic characteristic. (Mg0.6Zn0.4)1−yNiyTiO3 (y = 0.01–0.2) compounds were made from mixed oxides stoichiometrically and ground in distilled water for 24 h in a ball mill with agate balls. The mixed solution was dried in the oven and calcined at 1100 °C for 4 h in a high-temperature furnace. A 3.5 wt% of a 12% PVA solution as a binder (Polyvinyl alcohol 500, Showa) was added into the calcined powder, granulated by sieving through a 100 mesh, and pressed into pellets, 11 mm in diameter and 5 mm in thickness, under 200 MPa pressure. The pellets were sintered at temperatures ranging from 1125 to 1250 °C for 4 h in the atmosphere. The heating and cooling rates of a high-temperature furnace were both set at 10 °C/min to obtain high-quality samples.
The crystalline phases of the calcined powder and the sintered ceramics were distinguished by X-ray diffraction (XRD) using CuKa (λ = 0.15406 nm) radiation with a Siemens D5000 diffractometer (Siemens, Berlin/Munich, Germany). Scanning electron microscopy (SEM; Philips, Amsterdam, The Netherlands) was utilized to observe the sintered surface’s microstructure, and an energy-dispersive X-ray spectrometer (Philips) was used to identify the existence of secondary phases. Calculation of lattice parameter was performed with the Rietveld method to fit at least five peaks from measured X-Ray patterns. The apparent densities were measured using the Archimedes method with the three sintered samples under the same process and average measured data to calculate relative densities. The permittivity (εr) and the quality factor (Q) at microwave frequencies were measured by the Hakki–Coleman dielectric resonator method [28,29]. This method utilizes parallel conducting plates and coaxial probes in TE011 mode, where TE represented transverse electric waves, the first two subscript integers denote the waveguide mode, and the third integer subscript indicates the order of resonance in an increasing set of discrete resonant lengths. The measurement system was connected to Anritsu network analyzer with model MS46122B. The measurement technique for the temperature coefficient of resonance frequency (τf) is almost the same as that of the quality factor measurement except for using a thermostat. The τf at microwave frequencies was measured, and note change of resonance frequency from 30 °C to 80 °C. The following formula can calculate the τf value (ppm/°C),
τ f =   f 2 f 1 f 1 T 2 T 1
where f1 and f2 mean the resonance frequencies at T1 = 25 °C and T2 = 80 °C, respectively.

3. Results and Discussion

Figure 1 shows the room-temperature XRD analysis of (Mg0.6Zn0.4)1−yNiyTiO3 y = 0.01–0.2 sintered at 1200 °C for 4 h. The (Mg0.6Zn0.4)0.95Ni0.05TiO3 with an ilmenite-type structure, the same as (Mg0.6Zn0.4)TiO3: trigonal (ICDD #01-073-7752), were identified as the main phase associated with the apparent second phase (Mg0.6Zn0.4)0.95Ni0.05Ti2O5 identical to MgTi2O5 (JCPDS File No. 82-1125). MgTi2O5, called the second phase, is typically formed as an intermediate phase during the crystal growth of what is challenging to eliminate from the sample when MgO and TiO2 react in a 1:2 molar ratio [30,31,32].
The formation of (Mg0.6Zn0.4)0.95Ni0.05Ti2O5 might degrade the microwave dielectric performance of (Mg0.6Zn0.4)0.95Ni0.05TiO3 ceramics. As expected, the second phase (Mg0.6Zn0.4)0.95Ni0.05TiO3 was enhanced at high temperatures since the Mg0.95Ni0.05Ti2O5 requires a high sintering temperature of 1450 °C [26,33]. In other words, less (Mg0.6Zn0.4)0.95Ni0.05Ti2O5 phase would exist at lower sintering temperatures [34]. Similar XRD analysis was identified for the samples with y = 0.05 sintering at 1125–1250 °C for 4 h (Figure 2).
The lattice parameters of (Mg0.6Zn0.4)1−yNiyTiO3 ceramics sintered at 1200 °C for 4 h with various y values were also calculated for further structure investigation and forming of solid solution (Table 1). A decrease in the lattice parameters was observed for (Mg0.6Zn0.4)1−yNiyTiO3 ceramics compared to that of (Mg0.6Zn0.4)TiO3 (ICDD #01-073-7752). The results explain that (Mg0.6Zn0.4)0.95Ni0.05TiO3 would form a solid solution due to the replacement of Mg2+, Zn2+ by 0.05 mole Ni2+. Formation of (Mg0.6Zn0.4)0.95Ni0.05TiO3 would result in a fluctuation in the lattice parameters because the smaller Ni2+ ions (radii = 0.069 nm) relative to the size of the Mg2+ ions (radii = 0.072 nm) and Zn2+ ions (0.082 nm) are added in (Mg0.6Zn0.4)TiO3, lead to the lattice of (Mg0.6Zn0.4)0.95Ni0.05TiO3 locally distorted in A-site. These data indicated that with the partial replacement of Mg2+, Zn2+ by 0.05 mole Ni 2+, formation of (Mg0.6Zn0.4)0.95Ni0.05TiO3 solid solution would demonstrate a decrease in the lattice parameters from (a = 0.5064, and c = 1.3918 nm) in (Mg0.6Zn0.4)TiO3 to (a = 0.5063, and c = 1.3912 nm).
The micrographs of (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at different temperatures for 4 h are revealed in Figure 3. The microstructures of samples were not dense at 1125 °C. Grain size gradually increased with temperature and became highly uniform at 1200 °C. The growth of rod-shaped grain was enhanced at temperatures higher than 1225 °C and possibly the second phase. EDS analysis for Spot A–D and E–F identified the quantity composition of (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at 1200 °C and (Mg0.6Zn0.4)0.95Ni0.05Ti2O5 sintered at 1250 °C for 4 h, as shown in Figure 4a,b, respectively. Two kinds of grains possessed Mg, Zn, Ni, Ti, and O ions with different stoichiometric compositions (Mg0.6Zn0.4)0.95Ni0.05TiO3 and (Mg0.6Zn0.4)0.95Ni0.05Ti2O5, respectively, were clearly distinguished in the multiphase ceramics. The grain composition analysis observed from points A–F demonstrated that (Mg0.6Zn0.4)0.95Ni0.05TiO3 and (Mg0.6Zn0.4)0.95Ni0.05Ti2O5 form solid solutions to a certain degree, which is consistent with XRD analysis.
The apparent and relative density of the (Mg0.6Zn0.4)1−yNiyTiO3 with y = 0.01–0.2 sintered at different temperatures (1125–1250 °C) for 4 h is demonstrated in Figure 5. The relative density of the (Mg0.6Zn0.4)1−yNiyTiO3 exceeded 91.6% of the theoretical density (TD) in all specimens. The apparent density of the samples increased when the temperature increased from 1125 to 1200 °C. This increase in the apparent density can be attributed to the formation of dense microstructures, as observed in Figure 3d. The apparent density decreased when the temperature exceeded 1200 °C.
This decrease in the apparent density could result from the formation of a porous microstructure with large pores, as observed in Figure 3e,f. The apparent density and its corresponding TD increased from 4.15 g/cm3 (94.32% TD) to its maximum value of 4.39 g/cm3 (99.77%TD), and then decreased to 4.33 g/cm3 (98.41%TD) for the (Mg0.6Zn0.4)0.95Ni0.05TiO3 ceramics.
The permittivity and quality factor (Qf) of (Mg0.6Zn0.4)1−yNiyTiO3 ceramics sintered at temperatures for 4 h with various y values (0.01–0.2) are shown in Figure 6. The relationship between permittivity (εr) values and temperature reveals the same trend between densities and temperature since higher density means lower porosity inside samples. The permittivity increased with increasing temperature, reaching a maximum of 19.3 at 1200 °C and after that decreasing. Thus, increasing sintering temperature does not always result in a higher permittivity (εr). The permittivity of the well-sintered (Mg0.6Zn0.4)0.95Ni0.05TiO3 ceramics ranged from 17.9 to 19.3 at 1125–1200 °C. A maximum εr value of 19.3 was reached for the (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at 1200 °C for 4 h.
The Qf value is a crucial index for microwave dielectric ceramics applications because a high Qf value represented a lower dielectric loss for microwave devices. Qf values increased when temperature increased from 1125 °C to 1200 °C (Figure 6), consistent with the variation of bulk density, as shown in Figure 5. The increase of Qf value at low temperatures is attributed to the rise in density identical to grain growth uniformity, as shown in Figure 3. A maximum Qf value of 165,000 (GHz) was observed for the (Mg0.6Zn0.4)0.95Ni0.05TiO3 ceramics sintered at 1200 °C for 4 h. Qf value reached a maximum at 1200 °C, and after that, it decreased. The Qf value degraded due to the inhomogeneous grain growth, which leads to a reduction of density, as mentioned in Figure 3. There are several possible mechanisms to cause microwave dielectric loss, such as the lattice vibrational modes, the pores, second phases, impurities, and lattice defects [35]. Furthermore, apparent density also closely affects the dielectric loss and has also been demonstrated in microwave dielectric materials [35,36]. The Qf value and the variation of density of the (Mg0.6Zn0.4)0.95Ni0.05TiO3 show a similar trend, and hence apparent density is one of the dominant factors to cause the dielectric loss in this ceramic system.
As illustrated in Figure 7, the temperature coefficient of resonance frequency (τf) of (Mg0.6Zn0.4)1−yNiyTiO3 at temperatures ranging from 1125 to 1250 °C with x = 0.01–0.2. τf values fluctuate slightly at the whole temperature range. As expected, the synthesized composition persisted unchanged, and hence no noticeable fluctuation of the τf value was shown. As we know, τf is positively related to the synthesized composition, the second phase, and the material’s additive. The measured τf values were from −62.9 to −66 ppm/°C as the (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at 1125–1250 °C. At 1200 °C, a τf of (Mg0.6Zn0.4)0.95Ni0.05TiO3 with 4 h sintering time was measured around −65.4 ppm/°C. From the above discussion, microwave dielectric performance of (Mg0.6Zn0.4)1−yNiyTiO3 with various y value sintered at 1200 °C for 4 h was shown in Table 2 further demonstrated that best performance was found for y = 0.05. We also observed that Qf value is sensitive to Ni content, and however, permittivity and τf remain slightly fluctuate with varied Ni content. The sintering temperature affects the permittivity and Qf value to some extent. However, τf is not sensitive to sintering temperature. Table 3 revealed the comparison of the proposed dielectric with other similar documented dielectric ceramics. Among these three compositions, (Mg0.6Zn0.4)0.95Ni0.05TiO3 can be sintered under the lowest thermal budget (12.5% reduction as compared to MgTiO3) and measured highest Qf value (14.5% boost as compared to (Mg0.6Zn0.4)TiO3) with comparable εr and τf value. Overall, (Mg0.6Zn0.4)0.95Ni0.05TiO3 was demonstrated with excellent microwave performance and possible to manufacture high-performance substrates utilized in the microwave devices and circuits.
As mentioned previously, a dielectric resonator’s thermal stability is defined by the temperature coefficient of the resonance frequency. Dielectric materials systems with near-zero τf are crucial to fabricate substrates used in microwave frequency. In order to obtain near-zero τf dielectric material, SrTiO3 [34] possessed ultrahigh positive τfr ~205, Qf value ~4200 and τf ~ 1700 ppm/°C under 1350 °C) was chosen to mix with (Mg0.6Zn0.4)0.95Ni0.05TiO3f ~ −66 ppm/°C under 1200 °C). According to the mixture rules for τf, 0.94(Mg0.6Zn0.4)0.95Ni0.05TiO3 − 0.06SrTiO3 was fabricated and analyzed in advance. Figure 8 shows τf ranging from +46.6 to 68.5 ppm/°C under 1125–1225 °C. Thus, we believe a near-zero τf dielectric material with excellent microwave performance will be realized possibly after stoichiometry calculation and experiments in future work.

4. Conclusions

The performance of the (Mg0.6Zn0.4)1−yNiyTiO3 at microwave frequency was investigated and analyzed. The dielectric ceramics with an ilmenite-type structure belonging to the trigonal space group R-3 could be founded in all samples. With the substitution of Ni2+ for Mg2+ and Zn2+, the densification of (Mg0.6Zn0.4)1−yNiyTiO3 was significantly enhanced. Compared to documented dielectric ceramics, excellent microwave performance can be demonstrated for (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at 1200 °C/4h with εr~19.3, Qf value ~165,000 (GHz) and τf ~ −65.4 ppm/°C. τf of 0.94(Mg0.6Zn0.4)0.95Ni0.05TiO3 − 0.06SrTiO3 shows ~ +68.5 ppm/°C and, thus, a near-zero mixture could be obtained in future work. The (Mg0.6Zn0.4)0.95Ni0.05TiO3 microwave dielectric material system shows ultrahigh potential for applications in devices and systems of satellite and wireless communications.

Author Contributions

Conceptualization, C.-H.S.; methodology, S.-H.L.; validation, C.-H.S. and C.-L.P.; formal analysis, C.-H.S.; investigation, S.-H.L.; data curation, C.-C.H. and C.-L.P.; writing—original draft preparation, C.-H.S. and S.-H.L.; writing—review and editing, S.-H.L.; visualization, S.-H.L.; supervision, S.-H.L.; project administration, S.-H.L.; funding acquisition, S.-H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Technology, Taiwan, under Grant No. MOST 108-2221-E-224-050, MOST 109-2622-E-224-013, and industrial cooperation with Live Strong Optoelectronics under contract no. Yuntech 109-3019-1.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are contained within the article.

Acknowledgments

The authors acknowledge the technical support from the Advanced Instrumentation Center of National Yunlin University of Science and Technology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sebastian, M.T.; Ubic, R.; Jantunen, H. Low-loss dielectric ceramic materials and their properties. Int. Mater. Rev. 2015, 60, 392–412. [Google Scholar] [CrossRef]
  2. Itaalit, B.; Mouyane, M.; Bernard, J.; Womes, M.; Houivet, D. Effect of Post-Annealing on the Microstructure and Microwave Dielectric Properties of Ba(Co0.7Zn0.3)1/3Nb2/3O3 Ceramics. Appl. Sci. 2016, 6, 2. [Google Scholar] [CrossRef]
  3. Guha, D.; Gajera, H.; Kumar, C. Cross-Polarized Radiation in a Cylindrical Dielectric Resonator Antenna: Identification of Source, Experimental Proof, and Its Suppression. IEEE Trans. Antennas Propag. 2015, 63, 1863–1867. [Google Scholar] [CrossRef]
  4. Achouri, K.; Yahyaoui, A.; Gupta, S.; Rmili, H.; Caloz, C. Dielectric resonator metasurface for dispersion engineering. IEEE Trans. Antennas. Propag. 2016, 62, 673–680. [Google Scholar] [CrossRef]
  5. Weng, M.H.; Liauh, C.T.; Lin, S.M.; Wang, H.H.; Yang, R.Y. Sintering Behaviors Microstructure and Microwave Dielectric Properties of CaTiO3-LaAlO3 Ceramics Using CuO/B2O3 Additions. Materials 2019, 12, 4187. [Google Scholar] [CrossRef] [Green Version]
  6. Belous, A.; Ovchar, O.; Jancar, B.; Spreitzer, M.; Annino, G.; Grebennikov, D.; Mascher, P. The Effect of Chemical Composition on the Structure and Dielectric Properties of the Columbites A2+Nb2O6. J. Electrochem. Soc. 2009, 156, 206–212. [Google Scholar] [CrossRef]
  7. George, S.; Sebastian, M.T. The Effect of Added Sintering Aids on Microwave Dielectric Properties of Li2(Zn0.9Ca0.1)Ti3O8 Ceramics and Applications. J. Eur. Ceram. Soc. 2010, 30, 2585–2592. [Google Scholar] [CrossRef]
  8. Shen, C.H.; Pan, C.L.; Lin, S.-H. A Study of the Effect of Sintering Conditions of Mg0.95Ni0.05Ti3 on Its Physical and Dielectric Properties. Molecules 2020, 25, 5988. [Google Scholar] [CrossRef] [PubMed]
  9. Zhang, M.; Li, L.; Xia, W.; Liao, Q. Structure and properties analysis for MgTiO3 and (Mg0.97M0.03)TiO3 (M = Ni, Zn, Co and Mn) microwave dielectric materials. J. Alloy. Compd. 2012, 537, 76–79. [Google Scholar] [CrossRef]
  10. Wakino, K. Recent development of dielectric resonator materials and filters in Japan. Ferroelectrics 1989, 91, 69–86. [Google Scholar] [CrossRef]
  11. Kim, H.T.; Byun, J.D.; Kim, Y. Microstructure and Microwave Dielectric Properties of Modified Zinc Titanates (I). Mater. Res. Bull. 1998, 33, 963–973. [Google Scholar] [CrossRef]
  12. Kim, H.T.; Nahm, S.; Byun, J.D.; Kim, Y. Low-Fired (Zn, Mg)TiO3 Microwave Dielectrics. J. Am. Ceram. Soc. 1999, 82, 3476–3480. [Google Scholar] [CrossRef]
  13. Sohn, J.H.; Inaguma, Y.; Yoon, S.O.; Itoh, M.; Nakamura, T.; Yoon, S.J.; Kim, H.J. Microwave Dielectric Characteristics of Ilmenite-Type Titanates with High Q Values. Jpn, J. Appl. Phys. 1994, 33, 5466–5470. [Google Scholar] [CrossRef]
  14. Jo, H.J.; Kim, J.S.; Kim, E.S. Microwave dielectric properties of MgTiO3-based ceramics. Ceram. Int. 2015, 41, 530–553. [Google Scholar] [CrossRef]
  15. Huang, C.L.; Shen, C.H. Dielectric Properties and Applications of Low Loss (1-x)(Mg0.95Co0.05)TiO3-xCa0.8Sm0.4/3TiO3 Ceramic System at Microwave Frequency. J. Alloy. Comp. 2009, 468, 516–521. [Google Scholar] [CrossRef]
  16. Tang, B.; Xiang, Q.; Fang, Z.; Zhang, X.; Xiong, Z.; Lid, H.; Yuan, C.; Zhang, S. Influence of Cr3+ substitution for Mg2+ on the crystal structure and microwave dielectric properties of CaMg1-xCr2x/3Si2O6 ceramics. Ceram. Int. 2019, 45, 11484–11490. [Google Scholar] [CrossRef]
  17. Shen, C.H.; Huang, C.L. Microwave dielectric properties and sintering behaviors of (Mg0.95Ni0.05)TiO3-CaTiO3 ceramic system. J. Alloy. Compd. 2009, 472, 451–455. [Google Scholar] [CrossRef]
  18. Chen, Y.B.; Tseng, Z.L.; Chen, L.C.; Lin, C.C.; Miao, H.Y.; Liu, J.H.; Lin, S.H. Crystal structure and microwave dielectric properties of [(Mg0.6Zn0.4)0.95Co0.05]2TiO4-modified Ca0.6La0.8/3TiO3 cordierite ceramics with a near-zero temperature coefficient. J. Mater. Sci. Mater. Electron. 2018, 29, 10709–10714. [Google Scholar] [CrossRef]
  19. Huang, C.L.; Tseng, Y.W. A low-loss dielectric using CaTiO3-modified Mg1.8Ti1.1O4 ceramics for applications in dielectric resonator antenna. IEEE Trans. Dielectr. Electr. Insul. 2014, 21, 2293–2300. [Google Scholar] [CrossRef]
  20. Huang, C.L.; Tseng, C.F.; Chen, Y.B.; Cheng, Y.C. New dielectric material system of (Mg0.95Zn0.05)TiO3–Ca0.61Nd0.26TiO3 at microwave frequency. J. Alloys Compd. 2008, 453, 337–340. [Google Scholar] [CrossRef]
  21. Lin, S.H.; Chen, Y.B. Structure and characterization of B2O3 modified yNd(Mg1/2Ti1/2)O3-(1−y)Ca0.8Sr0.2TiO3 ceramics with a near-zero temperature coefficient at microwave frequency. Ceram. Int. 2017, 43, 2368–2371. [Google Scholar] [CrossRef]
  22. Yang, W.R.; Huang, C.L.; Chen, Y.R. Microwave and Optical Technology Letters, John Wiley. In Proceedings of the IEEE APMC, Yokohama, Japan, 12–15 December 2006; pp. 1398–1401. [Google Scholar]
  23. Fiedziuszko, S.J. Dielectric materials, devices, and circuits. IEEE Trans Microw. Theory Tech. 2002, 50, 706–720. [Google Scholar] [CrossRef]
  24. Nakagoshi, Y.; Suzuki, Y. Dimensional change behavior of porous MgTi2O5 in reactive sintering. Ceram. Int. 2017, 43, 5541–5546. [Google Scholar] [CrossRef]
  25. Suzuki, Y.; Suzuki, T.S.; Shinoda, Y.; Yoshida, K. Uniformly porous MgTi2O5 with narrow pore-size distribution: XAFS Study, Improved in situ synthesis, and new in situ surface coating. Adv. Eng. Mater. 2012, 14, 1134–1138. [Google Scholar] [CrossRef]
  26. Shen, C.H.; Huang, C.L. Phase Evolution and Dielectric Properties of (Mg0.95M2+0.05)Ti2O5 (M2+ = Co, Ni, and Zn) Ceramics at Microwave Frequencies. J. Am. Ceram. Soc. 2009, 92, 384–388. [Google Scholar]
  27. Kuszyk, J.A.; Bradt, R.C. Influence of Grain Size on Effects of Thermal Expansion Anisotropy in MgTi2O5. J. Am. Ceram. Soc. 1973, 56, 420–423. [Google Scholar] [CrossRef]
  28. Hakki, B.W.; Coleman, P.D. A Dielectric Resonator Method of Measuring Inductive Capacities in the Millimeter Range. IEEE Trans. Microw. Theory Tech. 1960, 8, 402–410. [Google Scholar] [CrossRef]
  29. Courtney, W.E. Analysis and Evaluation of a Method of Measuring the Complex Permittivity and Permeability Microwave Insulators. IEEE Trans. Microw. Theory Tech. 1970, 18, 476–485. [Google Scholar] [CrossRef]
  30. Liao, J.; Senna, M. Crystallization of titania and magnesium titanate from mechanically activated Mg(OH)2 and TiO2 gel mixture. Mater. Res. Bull. 1995, 30, 385–392. [Google Scholar] [CrossRef]
  31. Huang, C.L.; Pan, C.L. Low Temperature Sintering and Microwave Dielectric Properties of (1-x)MgTiO3-xCaTiO3 Ceramics Using Bismuth Additions. Jpn. J. Appl. Phys. 2002, 41, 707–711. [Google Scholar] [CrossRef]
  32. Nakagoshi, Y.; Sato, J.; Morimoto, M.; Suzuki, Y. Near-zero volume-shrinkage in reactive sintering of porous MgTi2O5 with pseudobrookite-type structure. Ceram. Int. 2016, 42, 9139–9144. [Google Scholar] [CrossRef] [Green Version]
  33. Lin, S.H.; Chen, Y.B. Crystal structure and microwave dielectric properties of [(Mg0.6Zn0.4)0.95Co0.05]2TiO4 modified Ca0.8Sm0.4/3TiO3 ceramics. Ceram. Int. 2017, 43, 296–300. [Google Scholar] [CrossRef]
  34. Silverman, B.D. Microwave Absorption in Cubic Strontium Titanate. Phys. Rev. 1962, 125, 1921. [Google Scholar] [CrossRef]
  35. 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. 1997, 80, 1885–1888. [Google Scholar] [CrossRef]
  36. Cho, W.W.; Kakimoto, K.; Sato, H.O. High-Q Microwave Dielectric SrTiO3-Doped MgTiO3 Materials with Near-Zero Temperature Coefficient of Resonant Frequency. Jpn. J. Appl. Phys. 2004, 43, 6221–6224. [Google Scholar] [CrossRef]
Figure 1. XRD analysis of (Mg0.6Zn0.4)1−yNiyTiO3 (y = 0.01–0.2) sintered at 1200 °C for 4 h. +: (Mg0.6Zn0.4)1−yNiyTiO3, O: (Mg0.6Zn0.4)1−yNiyTi2O5).
Figure 1. XRD analysis of (Mg0.6Zn0.4)1−yNiyTiO3 (y = 0.01–0.2) sintered at 1200 °C for 4 h. +: (Mg0.6Zn0.4)1−yNiyTiO3, O: (Mg0.6Zn0.4)1−yNiyTi2O5).
Applsci 11 02952 g001
Figure 2. XRD analysis of (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at 1125–1250 °C for 4 h. [+: (Mg0.6Zn0.4)0.95Ni0.05TiO3, O: (Mg0.6Zn0.4)0.95Ni0.05Ti2O5].
Figure 2. XRD analysis of (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at 1125–1250 °C for 4 h. [+: (Mg0.6Zn0.4)0.95Ni0.05TiO3, O: (Mg0.6Zn0.4)0.95Ni0.05Ti2O5].
Applsci 11 02952 g002
Figure 3. SEM photographs of (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at temperature range (a) 1125, (b) 1150, (c) 1175, (d) 1200, (e) 1225 and (f) 1250 °C for 4 h.
Figure 3. SEM photographs of (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at temperature range (a) 1125, (b) 1150, (c) 1175, (d) 1200, (e) 1225 and (f) 1250 °C for 4 h.
Applsci 11 02952 g003
Figure 4. The EDS analysis of (a) (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at 1200 °C (b) (Mg0.6Zn0.4)0.95Ni0.05Ti2O5 sintered at 1250 °C for 4 h.
Figure 4. The EDS analysis of (a) (Mg0.6Zn0.4)0.95Ni0.05TiO3 sintered at 1200 °C (b) (Mg0.6Zn0.4)0.95Ni0.05Ti2O5 sintered at 1250 °C for 4 h.
Applsci 11 02952 g004
Figure 5. Dependent relationship of apparent density(g/cm3) and relative density (%) of the (Mg0.6Zn0.4)1−yNiyTiO3 ceramics with y = 0.01–0.2 under 1125–1250 °C.
Figure 5. Dependent relationship of apparent density(g/cm3) and relative density (%) of the (Mg0.6Zn0.4)1−yNiyTiO3 ceramics with y = 0.01–0.2 under 1125–1250 °C.
Applsci 11 02952 g005
Figure 6. Permittivity and quality factor of the (Mg0.6Zn0.4)1−yNiyTiO3 with y = 0.01–0.2 under 1125–1250 °C.
Figure 6. Permittivity and quality factor of the (Mg0.6Zn0.4)1−yNiyTiO3 with y = 0.01–0.2 under 1125–1250 °C.
Applsci 11 02952 g006
Figure 7. τf value V.S sintering temperature of (Mg0.6Zn0.4)1−yNiyTiO3 with y = 0.01–0.2.
Figure 7. τf value V.S sintering temperature of (Mg0.6Zn0.4)1−yNiyTiO3 with y = 0.01–0.2.
Applsci 11 02952 g007
Figure 8. Microwave performance of 0.94(Mg0.6Zn0.4)0.95Ni0.05TiO3 − 0.06SrTiO3 mixture.
Figure 8. Microwave performance of 0.94(Mg0.6Zn0.4)0.95Ni0.05TiO3 − 0.06SrTiO3 mixture.
Applsci 11 02952 g008
Table 1. The calculated lattice parameter of (Mg0.6Zn0.4)1−yNiyTiO3 ceramic with y = 0.01–0.2.
Table 1. The calculated lattice parameter of (Mg0.6Zn0.4)1−yNiyTiO3 ceramic with y = 0.01–0.2.
ySinteringa (nm)c (nm)
ValueTemperature (°C)
0.0111750.50656 ± 0.000511.39113 ± 0.00082
12000.50702 ± 0.009001.39242 ± 0.00146
12250.50481 ± 0.000551.38946 ± 0.00089
0.0311750.50650 ± 0.000791.39209 ± 0.00170
12000.50713 ± 0.000571.39041 ± 0.00121
12250.50689 ± 0.000541.39147 ± 0.00087
0.0511750.50671 ± 0.000571.39158 ± 0.00122
12000.50670 ± 0.000321.39196 ± 0.00069
12250.50545 ± 0.000631.39144 ± 0.00103
0.0711750.50632 ± 0.000451.39188 ± 0.00096
12000.50585 ± 0.001061.38861 ± 0.00227
12250.50565 ± 0.000391.39175 ± 0.00064
0.112000.50577 ± 0.000681.39031 ± 0.00110
0.212000.50593 ± 0.000901.38992 ± 0.00146
Table 2. Microwave dielectric performance of (Mg0.6Zn0.4)1−yNiyTiO3 with various y values sintered at their optimized temperature for 4 h.
Table 2. Microwave dielectric performance of (Mg0.6Zn0.4)1−yNiyTiO3 with various y values sintered at their optimized temperature for 4 h.
y ValueS.T.DensityεrQ fτf
0.011200 °C/4 h4.3518.7120,000−66.5
0.031200 °C/4 h4.3618.8140,000−67.4
0.051200 °C/4 h4.3919.3165,000−65.4
0.071200 °C/4 h4.3719155,000−66
0.11200 °C/4 h4.2718.9135,000−59.7
0.21200 °C/4 h4.2518.690,000−57.6
Table 3. Comparison of the proposed dielectric with other similar documented dielectric ceramics.
Table 3. Comparison of the proposed dielectric with other similar documented dielectric ceramics.
CompositionS.T.εrQ fτfRef
MgTiO31350 °C/4 h17160,000−51[10]
(Mg0.6Zn0.4)TiO31200 °C/4 h19.8144,000−66[12]
(Mg0.6Zn0.4)0.95Ni0.05TiO31200 °C/4 h19.3165,000−65.4This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shen, C.-H.; Pan, C.-L.; Lin, S.-H.; Ho, C.-C. Microwave Performance, Microstructure, and Crystallization of (Mg0.6Zn0.4)1−yNiyTiO3 Ilmenite Ceramics. Appl. Sci. 2021, 11, 2952. https://0-doi-org.brum.beds.ac.uk/10.3390/app11072952

AMA Style

Shen C-H, Pan C-L, Lin S-H, Ho C-C. Microwave Performance, Microstructure, and Crystallization of (Mg0.6Zn0.4)1−yNiyTiO3 Ilmenite Ceramics. Applied Sciences. 2021; 11(7):2952. https://0-doi-org.brum.beds.ac.uk/10.3390/app11072952

Chicago/Turabian Style

Shen, Chun-Hsu, Chung-Long Pan, Shih-Hung Lin, and Cheng-Che Ho. 2021. "Microwave Performance, Microstructure, and Crystallization of (Mg0.6Zn0.4)1−yNiyTiO3 Ilmenite Ceramics" Applied Sciences 11, no. 7: 2952. https://0-doi-org.brum.beds.ac.uk/10.3390/app11072952

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