Ceramic dielectric materials with high dielectric strength and mechanisms of their internal factors affecting dielectric strength are significantly valuable for industrial application, especially for selection of suitable dielectric materials for high-power microwave transmission devices and reliable power transmission. Pure magnesium oxide (MgO), a kind of ceramic dielectric material, possesses great application potential in high-power microwave transmission devices due to its high theoretical dielectric strength, low dielectric constant, and low dielectric loss properties, but its application is limited by high sintering temperature during preparation. This work presented the preparation of a new type of multiphase ceramics based on MgO, which was MgO-1%ZrO2-1%CaCO3-x%MnCO3 (MZCMx, x = 0, 0.25, 0.50, 1.00, 1.50, in molar), and their phase structures, morphological features, and dielectric properties were investigated. It was found that inclusion of ZrO2 and CaCO3 effectively inhibited excessive growth of MgO grains by formation of second phase, while addition of MnCO3 promoted the grain boundary diffusion process during the sintering process and reduced activation energy for the grain growth, resulting in a lower ceramic sintering temperature. Excellent performance, including high dielectric strength (Eb = 92.3 kV/mm) and quality factor (Q × f = 216642 GHz), simultaneously accompanying low dielectric loss (< 0.03%), low temperature coefficient of dielectric constant (20.3×10-6 ℃-1, 85 ℃) and resonance frequency (-12.54×10-6 ℃-1), was achieved in MZCM1.00 ceramics under a relatively low sintering temperature of 1350 ℃. This work offers an effective solution for selecting dielectric materials for high-power microwave transmission devices.
【AIGC One Sentence Reading】:本研究通过掺杂碳酸锰降低氧化镁陶瓷烧结温度,制得了高介电强度、低介电损耗的陶瓷材料,为高功率微波传输器件提供了优质的介质窗材料解决方案,具有实际应用价值。
【AIGC Short Abstract】:本研究通过传统固相反应法制备了掺杂锰的氧化镁基陶瓷,有效降低了烧结温度,提升了介电强度和品质因数,同时保持了低介电损耗和稳定的温度系数。这一成果为高功率微波传输器件的介质窗材料选择提供了新方案,具有重大的科学研究和工程应用价值。
Note: This section is automatically generated by AI . The website and platform operators shall not be liable for any commercial or legal consequences arising from your use of AI generated content on this website. Please be aware of this.
The trend towards high integration, density, and power in electronic circuits has led to more stringent requirements for the dielectric strength, mechanical strength, and thermal conductivity of dielectric materials. The research for dielectric materials with high and reliable dielectric strength and properties is of great significance for engineering applications in various fields such as dielectric capacitors[1-2], high-power microwave device components[3-4] and substrates[5]. In the past decade, great efforts have been made to study the dielectric strength of different materials. Polymers usually have low dielectric constants, dense microstructure, and amorphous state, which are the best choices for high dielectric strength[6-7]. While ceramic materials have become a research hotspot due to their excellent temperature stability, aging stability, and extended service life[8-9]. However, characteristics like high dielectric constant, porous and coarse grain structure make ceramics easier to breakdown at low electric field, which are usually lower than one-tenth of that of polymers[10]. Therefore, searching for ceramics with high dielectric strengths is of great importance. At present, chemical composition regulation, including ion doping to modify phase structure, construct defects, or regulate microstructure, has been extensively studied and proven to be simpler and more feasible[11-12].
For MgO ceramics, lower dielectric constant (~10), higher bandgap (nearly 7.77 eV) and simple rock salt crystal structure contribute to high dielectric strength as 232 kV/mm[13-14]. However, the actual value is usually much lower than the theoretical one, which requires the regulation of the non-intrinsic factors, such as the ceramic microstructure. Former studies have shown that the addition of 1%-1.5% (in molar) ZrO2 and CaCO3 is proved effective to inhibit abnormal grain growth of MgO and promote pore discharge, leading to high density and mechanical strength, which are beneficial to enhance the dielectric strength[15-16]. Therefore, ZrO2 reinforced MgO ceramics possess great potential in high-voltage microwave dielectric application. However, sintering temperature of the above-mentioned MgO-based ceramics is higher than 1500 ℃, which limits their application in microwave capacitors and insulators. Fortunately, previous researches have indicated that MnCO3 is effective in lowering the sintering temperature of ceramics and enhancing the density[17⇓-19]. Hence, MnCO3 is utilized here to reduce the sintering temperature of MgO-based ceramics, and their relevant application potential substrate materials are worth exploring.
In this work, MgO-1%ZrO2-1%CaCO3-x%MnCO3 (MZCMx, x = 0, 0.25, 0.50, 1.00, 1.50, in molar) ceramics were prepared, and ZrO2 and MnCO3 are used as reinforcing phase and sintering additive, respectively. The microstructure evolution and dielectric properties were studied and discussed in detail. The results show that the doping of MnCO3 significantly reduces the sintering temperature of MgO by ion substitution to activate the lattice and reduce the sintering activation energy, which affects and changes the sintering behaviour and dielectric properties of MgO.
1 Experimental
1.1 Material preparation
Light magnesium oxide (MgO, 99.650%), ZrO2 (99.980%), CaCO3 (99.657%) and MnCO3 (99.916%) were mixed and ball milled for 6 h with ethanol by a planetary ball mill, then the dried powders were pressed into ϕ13 mm × 2 mm pellets and ϕ13 mm × 10 mm cylinders at a pressure of 200 MPa. Finally, these pellets and cylinders were sintered at 1520 ℃ (for pure MgO ceramics), 1540 ℃ (for MZCM0 ceramics), 1400 ℃ (for MZCM0.25 and MZCM0.50 ceramics), and 1350 ℃ (for MZCM1.00 and MZCM1.50 ceramics) for 2 h, respectively. The sintered pellets were machined into ϕ10.0 mm × 0.5 mm and coated with silver electrode (ϕ8.0 mm) for capacitance and dielectric strength measurement. The cylinders were machined into ϕ10.0 mm × 5.5 mm samples by plane and outer edge processing for microwave performance test.
1.2 Material characterization
The chemical composition of ceramic powders was measured by an X-ray fluorescence spectrometer (Axios, PANalytical, Netherlands). The density of the sintered ceramic samples was tested by Archimedes method, and the crystal structure of the materials was analyzed by an X-ray diffractometer (XRD, PANalytical 9430070 99991 series). The morphology and chemical composition of the ceramic samples were characterized by a field emission scanning electron microscope (SEM, Magellan400, FEI, USA). The dielectric strength of the samples (at least 8 pieces for each composition) at room temperature was tested using high-voltage source (WISMAN, DL120P600, 0-120 kV). The temperature and frequency dependence of the dielectric constant and dielectric loss were evaluated via a broadband dielectric Novocontrol Alpha spectrometer (Novocontrol Technologies, Germany). Vector network analyzer (Agilent E8362B) and temperature chamber were used to measure the microwave dielectric properties of the samples. The quality factor (Q × f) of samples was estimated by the TE01δ mode dielectric resonator method. The resonant frequency temperature coefficient τf was calculated by the following equation:
where f85 and f25 represent the resonant frequencies at 85 and 25 ℃, respectively. ∆T represents the temperature difference, 60 ℃. The capacitance temperature coefficient τc could be calculated using a similar equation:
where C2 and C1 represent the capacitances at T2 and 25 ℃, respectively.
2 Results and discussion
Fig. 1 shows the dependence of densities of MZCMx (x = 0, 0.25, 0.50, 1.00, 1.50) ceramics versus x and sintering temperature. According to the changing trend of density, the optimal sintering temperature for sample densification can be obtained, which is 1520 ℃ for pure MgO, 1540 ℃ for x = 0, 1400 ℃ for x = 0.25, 0.50, and 1350 ℃ for x = 1.00, 1.50, respectively. The optimal sintering temperature decreases with the increase of x.
Figure 1.Densities of the pure MgO and MZCMx (x = 0, 0.25, 0.50, 1.00, 1.50) ceramics at different sintering temperatures (1300-1560 ℃, 2 h)
Table 1. Microwave dielectric properties of pure MgO and MZCMx ceramics
Sample
τc/(×10-6, ℃-1, -25 ℃)
τc/(×10-6, ℃-1, 85 ℃)
(Q × f)/GHz
τf/(×10-6,℃-1)
Pure MgO
79.6
85.1
138240
-35.62
MZCM0
63.9
74.4
124623
-32.36
MZCM0.25
28.6
34.4
130330
-20.64
MZCM0.50
20.1
46.1
156196
-18.36
MZCM1.00
4.91
20.3
216642
-12.54
MZCM1.50
16.06
78.4
123216
-14.60
where G0 and Gt represent the initial grain size as well as the grain size after sintering for t time at T temperature, R is the ideal gas constant, n and Q represent the grain growth index and the grain growth activation energy, respectively. To solve these two key parameters, the above equation can be deformed:
where G is the average grain size of the ceramic, and k0 is the constant. The grain growth index n can be solved using the graphical method by controlling the sintering temperature and sintering time of the ceramics to make the logarithmic fitting curves of the average grain size to the sintering time and temperature, respectively[20].
The calculation results of pure MgO ceramics show that the grain growth index n = 3 of pure MgO ceramics corresponds to the volume diffusion process, and the grain growth activation energy Q is about 556.9 kJ/mol[20]. Fig. 2 is the solution of grain growth index and grain growth activation energy of MZCM1.00 ceramics. The grain growth index n of MZCM1.00 ceramics sintered at 1300 ℃ is 1.97 (≈ 2). Among them, the size of n represents different diffusion mechanisms in the sintering process of ceramics. n = 2 corresponds to the grain boundary diffusion process, while the grain boundary diffusion mode can be regarded as a fast channel for material diffusion. Therefore, MZCM1.00 ceramic grain growth gradually changes from the grain boundary diffusion mode to the intrinsic bulk diffusion mode of MgO with the increase of temperature, implying that the introduction of Mn significantly promotes the grain growth process during the initial stage of ceramic sintering. Meanwhile, due to the change of grain growth index during the sintering process of ceramics, n = 2.5 is taken to further solve the grain growth activation energy. The average grain growth activation energy is about 301.56 kJ/mol, which is much lower than that of pure MgO ceramics, further proving the promotion effect of Mn on the sintering process of MgO ceramics.
Figure 2.Kinetic curves of grain growth in MZCM1.00 ceramics
Fig. 3 shows the XRD patterns of MZCMx ceramics. A composite structure composed of cubic-MgO (PDF#45-0946), cubic-ZrO2 (PDF#49-1642) and glass phase can be clearly observed while the diffraction peaks of cubic-ZrO2 weaken with the increase of MnCO3 content, which means the decrease of cubic-ZrO2 content. Firstly, Ca2+ solid dissolves into MgO lattice completely, and then the introduction of MnCO3 results in the replacement of Ca2+ and the formation of glass phase.
Figure 3.XRD patterns of pure MgO and MZCMx (x = 0, 0.25, 0.50, 1.00, 1.50) ceramics
Fig. 4 reveals the morphological evolution and the grain size distribution of different thermal etched MZCMx samples. Initially, the average grain size of pure MgO ceramics is high up to 22.94 μm, and the addition of ZrO2 and CaCO3 forms the nano-sized second phase at the grain boundaries of MgO, leading to the grain size decreasing to 5.95 μm. After that, with the doping of MnCO3, the content of the second phase decreases, resulting in an increase for grain size, but the overall size is still smaller than that of pure MgO. The ion vacancy concentration significantly affects the grain growth and sintering process. Similar condition happens to Cooper et al.[21], who found that increasing the concentration of U4+ vacancies could significantly promote the growth of UO2 grains. The sintering process of MgO is dominated by anion diffusion, and the doping of MnCO3 further reduces the oxygen vacancy concentration and inhibits the mass transfer process during the later stages of sintering, resulting in an increase in grain size.
Figure 4.FESEM images of thermal etched surface morphologies of ceramics with insets showing grain size distribution of ceramics
X-ray photoelectron spectroscopy (XPS) is used to characterize the changes in oxygen vacancy concentration, and MZCM1.00 and MZCM1.50 ceramic samples are selected to test due to the limited doping concentration of MnCO3. The samples were subjected to argon ion etching to eliminate adsorbed oxygen on the sample surfaces prior to testing for the oxygen element, and the results of the fitted analyses are shown in Fig. 5. Oxygen in the samples mainly exists in the forms of lattice oxygen (LO) and oxygen vacancy (OV), and doping of Mn ions leads to a reduction of the ratio of the vacancy oxygen peak area to the lattice oxygen peak area, implying that the concentration of oxygen vacancies decreases in the systems. It is worth noting that the oxygen vacancies in the ceramic systems prepared in this work originate from intrinsic oxygen vacancies due to the susceptibility of MgO to Schottky defect reactions[22].
The Weibull distribution is widely used to describe the dielectric strength[12]. In this model, the probability of dielectric breakdown is defined as follows:
where Ei, Eb and β are specific breakdown strength of single sample, characteristic breakdown strength and shape parameter that indicates the width of distribution, respectively. Fig. 6(a) illustrates the Weibull distribution of the dielectric strength for MZCMx ceramics, in which Xi and Yi are two parameters which can be described as follows:
where i and n are sample sequence in specific order and total amount of samples for each composition. The fitting dielectric strengths of MZCMx are 113.95, 99.4, 96.2, 92.3, and 85.2 kV/mm for x = 0, 0.25, 0.50, 1.00, and 1.50, respectively. The fine grain and dense structure of MZCM0 ceramics allow for a significant increase in dielectric strength. And doping of MnCO3 results in reduction in dielectric strength due to the weakened second phase and intergranular bonding strength[23]. Fig. 6(b) shows the temperature dependence of the dielectric constant and dielectric loss of MgO-based dielectric ceramics at 1 MHz, respectively. The dielectric losses of all samples are less than 0.1% in the tested temperature range, presenting a low loss characteristic. In addition, the dielectric properties of all samples maintain good stability in the temperature range of 50-300 ℃, which is beneficial to their application as core materials in extreme environments.
Figure 6.Dieletric properties of pure MgO and MZCMx (x = 0, 0.25, 0.50, 1.00, 1.50) ceramics
The dielectric properties of MgO-based ceramics are characterized in this work, which are shown in Table 1. MZCM1.00 ceramics exhibit low τc in the temperature range of -25-85 ℃, which further proves the good temperature stability. The microwave dielectric properties of the samples were measured at a resonant frequency range of 10.2-10.5 GHz. In particular, MZCM1.00 ceramics have a Q × f of up to 216642 GHz, and τf is relatively close to zero (-12.54×10-6 ℃-1), which is conducive to the integration and miniaturization of electronic circuits[24]. The addition of MnCO3 reduces oxygen vacancies, resulting in the creation of free electrons. Therefore, the overabundant MnCO3 causes excessive electrons, increasing dielectric loss and decreasing Q × f [25].
3 Conclusions
In summary, MnCO3 doped MgO-ZrO2-CaCO3 high- voltage dielectric ceramics are designed and prepared. The ultrahigh dielectric strength of 92.3 kV/mm was achieved in MZCM1.00 ceramics, and the sintering temperature is significantly reduced to 1350 ℃. The remarkable improvement of dielectric strength is mainly due to two main factors. Firstly, MgO has the potential to achieve high dielectric strength. Secondly, the introduction of ZrO2 and CaCO3 modifies the microstructure, eliminates external factors, and fully utilizes its dielectric properties. At the same time, MnCO3 can reduce the sintering temperature of MgO ceramics by accelerating mass transfer caused by ionic vacancies. The series of ceramics with high dielectric strength and medium sintering temperature also show high Q × f, which makes the MgO-based ceramic a candidate material for microwave applications. This MgO-based ceramic also has the advantages of simple components and preparation process, displaying the great potential for the application in dielectric materials of high-power microwave transmission devices.
[1] CAI Z M, FENG P Z, ZHU C Q et al. Dielectric breakdown behavior of ferroelectric ceramics: the role of pores. J. Eur. Ceram. Soc., 2533(2021).
[2] PAN H, LI F, LIU Y et al. Ultrahigh-energy density leadfree dielectric films via polymorphic nanodomain design. Science, 578(2019).
[3] CHEN K B, ZHOU X D, GAO M. Research progress and application of high power microwave technology. Winged Missile Journal, 1(2019).
[4] CHAI Y Y. Research of the Ku band compact high-power microwave output window, 1-4(2016).
[5] ZHANG X, WANG T, YU Q Q et al. Research progress of high-power waveguide window. High Power Laser and Particle Beams, 023001(2021).
[6] LI Q, CHEN L, GADINSKI M R et al. Flexible high-temperature dielectric materials from polymer nanocomposites. Nature, 576(2015).
[8] HUANG Y, CHEN Y, LI X et al. Enhanced dielectric breakdown strength in Ni2O3 modified Al2O3-SiO2-TiO2 based dielectric ceramics. J. Eur. Ceram. Soc., 3861(2018).
[9] LIU M, CAO M, ZENG F et al. Fine-grained silica-coated barium strontium titanate ceramics with high energy storage. Ceram. Inter., 20239(2018).
[10] PING W W, LIU W F, LI S T. Enhanced energy storage property in glass-added Ba(Zr0.2Ti0.8)O3-0.15(Ba0.7Ca0.3)TiO3 ceramics and the charge relaxation. Ceram. Inter., 11388(2019).
[11] ZHANG J, WANG J, GAO D et al. Enhanced energy storage performances of CaTiO3-based ceramic through A-site Sm3+ doping and A-site vacancy. J. Eur. Ceram. Soc., 352(2021).
[13] ROESSLER D M, WALKER W C. Electronic spectrum and ultraviolet optical properties of crystalline MgO. Phys. Rev., 733(1967).
[14] BEAUCHAMP E K. Effect of microstructure on pulse electrical strength of MgO. J. Amer. Ceram. Soc., 484(1971).
[15] GÓMEZ-RODRÍGUEZ C, GARCÍA-QUIÑONEZ L V, AGUILAR-MARTÍNEZ J A et al. MgO-ZrO2 ceramic composites for silicomanganese production. Materials, 2421(2022).
[16] ŚNIEŻEK E, SZCZERBAA J, STOCH P et al. Structural properties of MgO-ZrO2 ceramics obtained by conventional sintering, arc melting and field assisted sintering technique. Mater. Desi., 412(2016).
[17] WANG X, LIANG P, CHAO X et al. Dielectric properties and impedance spectroscopy of MnCO3-modified (Ba0.85Ca0.15)(Zr0.1Ti0.9)O3 lead-free ceramics. J. Am. Ceram. Soc., 1506(2015).
[18] CEN Z, DONG Z, XU Z et al. Improving fatigue properties, temperature stability and piezoelectric properties of KNN-based ceramics via sintering in reducing atmosphere. J. Eur. Ceram. Soc., 4462(2021).
[19] WANG T, LI Z. Effects of MnO2 addition on the structure and electrical properties of PIN-PZN-PT ceramics with MPB composition. J. Mater. Sci.: Mater. Elect., 22740(2020).
[20] HUANG Q Z, LU G M, SUN Z et al. Effect of TiO2 on sintering and grain growth kinetics of MgO from MgCl2·6H2O. Mater. Trans. B, 344(2013).
[21] COOPER M W D, STANEK C R, ANDERSSON D A. The role of dopant charge state on defect chemistry and grain growth of doped UO2. Act. Mater., 403(2018).
[23] ZHANG C, CHEN Y, LI X et al. Effect of LiF addition on sintering behavior and dielectric breakdown mechanism of MgO- based microwave dielectric ceramics. J. Mater., 478(2021).
[24] TAN Z, LIN H, SONG K et al. Effects of TiO2 additive on ultra-low-loss MgO-LiF microwave dielectric ceramics. Ceram. Inter., 5753(2022).
[25] XIONG Z, TANG B, ZHANG X et al. Suppression of Ti3+ generation in Ba3.75Nd9.5Ti17.5M0.5O54 (M = Cu, Cr, Al, Mn) ceramics. Ceram. Inter., 19058(2018).
Tools
Get Citation
Copy Citation Text
Zhixiang WANG, Ying CHEN, Qingyang PANG, Xin LI, Genshui WANG. Sintering Behaviour and Dielectric Properties of MnCO3-doped MgO-based Ceramics[J]. Journal of Inorganic Materials, 2025, 40(1): 97