Next Article in Journal
Achieving Tunable Microwave Absorbing Properties by Phase Control of NiCoMnSn Alloy Flakes
Next Article in Special Issue
Corrosion Behaviour of Cemented Carbides with Co- and Ni-Alloy Binders in the Presence of Abrasion
Previous Article in Journal
Analysis of the Effect of Magnetic Field on Solidification of Stainless Steel in Laser Surface Processing and Additive Manufacturing
Previous Article in Special Issue
Fe-Si Intermetallics/Al2O3 Composites Formed between Fe-20% Si and Fe-70.5% Si by SHS Metallurgy Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sr2+ Ion Substitution Enhanced Dielectric Properties of Co(2)Z Ferrites for VHF Antenna Substrate

1
School of Management and Economics, University of Electronic Science and Technology of China, Chengdu 610054, China
2
School of Mathematical Science, University of Electronic Science and Technology of China, Chengdu 610054, China
3
School of Information Science and Technology, Southwest Jiaotong University, Chengdu 611756, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(9), 1541; https://doi.org/10.3390/met12091541
Submission received: 6 July 2022 / Revised: 5 September 2022 / Accepted: 14 September 2022 / Published: 18 September 2022
(This article belongs to the Special Issue Metal-Ceramic Composites Fabricated by Powder Metallurgy Method)

Abstract

:
The effect of Sr2+ ions on the microstructure and high frequency properties of 2.5 wt% Bi2O3 added to Co(2)Z hexaferrites (3Ba(1-x)SrxO•2CoO•12Fe2O3, x = 0.0, 0.2, 0.4 and 0.6) synthesised using the solid-state reaction method was investigated. Experimental results indicate that the dielectric properties were markedly enhanced with the increase in the content of Sr2+ ions, thereby increasing the miniaturisation factor, which enables a size reduction in a long frequency range. Slight changes to saturation magnetisation (Ms) and coercivity (Hc) were observed, i.e., the saturation magnetisation (Ms) decreased from 39.99 to 38.11 emu/g, and coercivity (Hc) increased from 59.05 to 65.21 Oe when x increased from 0.0 to 0.6. Meanwhile, ε′ increased from approximately 8 to 12, indicating the invariability in μ′. In addition, the processed materials exhibit relatively low magnetic loss and dielectric loss (magnetic loss tanδμ ≈ 0.08 and dielectric loss (tanδε ≈ 0.007)). These results indicate that the substituted CO(2)Z ferrites have excellent potential in high-frequency antenna applications.

1. Introduction

The rapid development of information and communication technology has led to greater demand on the size and properties of modern communication equipment. Lightweight, high-performance ferrite-based antennas play important roles in wireless communication systems. Another important factor is the miniaturisation of antennas, which has become the topic of recent research [1,2,3]. However, one problem is miniaturisation of the antenna negatively affects the performance of the wireless communication system. Thus, the search for a range of new approaches to determine size and performance is urgent. One method to achieve miniaturisation is to increase the miniaturisation factor, which depends on the refractive index (n = (μ′ε′)1/2, where μ′ and ε′ are actual parts of permeability and permittivity, respectively). Therefore, n increases as the value of μ′, ε′ or both increases [4]. However, this approach is insufficient in the reception and transmission process of electromagnetic waves, as matching impedance between the antenna substrate and free space weakens the excitation of the surface wave, which is the key to mutual coupling between antenna arrays, causing deterioration of antenna radiation performance [5,6,7]. Therefore, adjusting the impedance of the antenna substrate to achieve impedance matching is important to ensuring the radiation performance of the antenna. The impedance (Z) of electronic materials is closely correlated to μ′ and ε′, according to the definition of Z in the following equation [8]:
Z = η0(μ′/ε′)1/2
where η0 is the impedance of free space. If μ′ and ε′ values could be tailored to become close or equal to each other, then Z becomes close to η0. In addition, low loss properties are equally important for the antenna substrate material, as loss control is of paramount importance to reducing energy consumption [9,10,11]. Therefore, ferrite materials are ideal for antenna substrate applications that possess μ′ and ε′, relatively high working frequency and low magnetic and dielectric loss.
The soft magnetic material Ba3Co2Fe2O41 (Co2Z) hexaferrite has been broadly used in high-frequency devices owing to its high intrinsic magnetic anisotropy field, moderate saturation magnetisation and permittivity [12,13,14]. It has high permeability and low loss properties in the range of terrestrial digital multimedia broadcasting (T-DMB) frequency [15], and its crystal structure consists mainly of stacked layers of tetrahedral and octahedral Fe3+ sites shared by Co2+ (3d7) ions. The high-temperature magnetoelectric coupling is ascribed to magnetic Co2+ (3d7) ions contributing to enhancing the electronic exchange strength [16]. Nanomaterials play important roles in various applications including energy, biomedical, sensing and pharmaceuticals [17,18,19]; recent research has focused on Co(2)Z barium ferrites to tailor the magnetic–dielectric properties through substituting Ba3+ ions using other ions by introducing nanomaterials. Amongst them, Sr2+ ion substitution is considered an effective method to adjust the magnetic–dielectric properties [20,21]. In addition, Bi2O3 sintering aids can not only lower sintering temperature but also tailor magnetic–dielectric properties of Co(2)Z barium ferrites. Harris et al. investigated how Bi2O3 aids modified M-type barium ferrites to achieve equal permeability and permittivity over a long frequency range. Our previous research explored the effect of ion substitution on the adjustment of magnetic properties of M-type barium ferrite, as well as Bi2O3 aids on the sintering temperature and ferrite densification [22,23]. Results showed that combined with the excellent tuneable characteristics of Co(2)Z barium ferrites, improving high-frequency electromagnetic characteristics by ion substitution is an effective method to achieve matching permeability and permittivity as well as high miniaturisation factor. Here, 3Ba(1-x)SrxO•2CoO•12Fe2O3 (x = 0.0, 0.2, 0.4 and 0.6) ferrites with 2.5wt% Bi2O3 were prepared. The magnetic–dielectric properties of the resultant materials were studied to achieve equal permeability and permittivity and low loss properties over a long frequency band in the high frequency range [24].

2. Experiment and Measurement

2.1. Materials and Methods

Sr2+ ion-substituted Co(2)Z barium ferrites 3Ba(1-x)SrxO•2CoO•12Fe2O3 with added 2.5 wt% Bi2O3 were synthesised using analytical grade BaCO3 (AR grade, ≥99%), Fe2O3 (AR grade, ≥99.5%), Co2O3 (AR grade, ≥99.5%), SrO (AR grade, ≥99%) and Bi2O3 (AR grade, ≥99.5%). Subsequently, the objective products were synthesised through the conventional solid-state reaction method. The raw powders were mixed in a ball mill for 10 h, with zirconia balls and deionised water as the milling media. Then, the mixed powders were dried and pre-sintered at 1150 °C for 4 h in a muffle furnace. Bi2O3 was added to the pre-sintered powder and then milled once more for 12 h. After drying, the milled powders were pelleted by adding 10 wt% polyvinyl alcohol (PVA) and then pressed into thick plates and rings. Finally, the dense samples were sintered at 925 °C for 4 h.

2.2. Measurements

The phase constitution of the ferrites was detected by an X-ray diffractometer (XRD, DX-2700, Haoyuan Co., Chengdu, China) with Cu-Kα radiation. The microtopography of the ferrite surface was detected using a scanning electron microscope (SEM, JEOL, JSM-6490, Tokyo, Japan). The complex permeability and dielectric constant were measured with a HP-4291BRF impedance analyser (Agilent, Santa Clara, CA, USA). The bulk density was measured using an auto density tester (GF-300D, AND Co., Tokyo, Japan) based on Archimedes’ principle. A vibrating sample magnetometer (VSM, EZ Model 10, MicroSense, Encinitas, CA, USA) was used to measure the magnetisation hysteresis loops. All measurements were performed at room temperature.

3. Results and Discussion

3.1. SEM

Surface topography of cross sections of the ferrite samples with different Sr2+ ion substitution x values is displayed in Figure 1. On the one hand, a marked change in the average grain size is observed, indicating that as the content of substituted Sr2+ ions increases, the average grain size shows a downward trend. The average grain size can be calculated as 1.03, 0.89, 0.67 and 0.61 μm for different x values using a statistical method with the following formula [25]:
G a = 1.5   L M   N
where L is the total line length, and M and N are the magnification and the total number of intercepts, respectively. The decrease in grain size with the increase in Sr2+ ion substitution is due to the smaller ion radius of Sr2+ ions than that of Ba2+; small ion radius always causes low grain size [26]. On the other hand, with the increase in Sr2+ ion substitution, more pores are observed due to the increased difficulty in crystallisation and clusters of grains [27].

3.2. XRD

The XRD patterns of the ferrites with different Sr2+ ion substitution x values are shown in Figure 2. Normal Co(2)Z barium ferrite phase can be obtained by adding 5 wt% Bi2O3 aids, from which the peaks with normal BaFe12O19 phase and BiFeO3 phase can be obtained, indicating no other phase or structure generated during the sintering process.
The formation of the BiFeO3 dielectric phase is due to the Bi3+ ions from superfluous Bi2O3 aids combining Fe3+ and O2− ions [28,29]. With more Sr2+ ion substitutions, the peak phase of the Co(2)Z ferrites shifts towards a higher angle direction, indicating that the lattice constant decreases with the increase in Sr2+ ion content, according to the relationship between the phase and lattice constant [30,31].

3.3. Magnetic and Dielectric Properties

Figure 3 shows the magnetic hysteresis loops and magnetic properties of samples sintered at five different temperature points (x = 0.0, 0.2, 0.4 and 0.6). The magnetic hysteresis loops in Figure 3a indicate that the Sr2+ ion-substituted barium ferrites have excellent soft magnetic properties at low temperature. The loops also indicate that the magnetisation slightly weakened as x increased from 0.0 to 0.6. The coercivity and saturation magnetisation value can be induced based on the loops. As shown in Figure 3b, the saturation magnetisation (Ms) decreased from 39.99 to 38.11 emu/g when x increased from 0.0 to 0.6. Meanwhile, coercivity (Hc) increased from 59.05 to 65.21 Oe when x increased from 0.0 to 0.6. The decrease in saturation magnetisation could be attributed to the decrease in grain size, according to the Neel theory of two sublattices [32], whereas the coercivity changes with inverse proportion to the saturation magnetisation. Their relationship is indicated by the following equation [33,34]:
M s = 0.96 K H c
where Ms is saturation magnetisation, K is a dependence constant, and Hc is coercivity. Frequently, larger grain size conducts lower coercivity due to domain wall pinning that requires high energy for switching [35].
Figure 4 shows the complex magnetic permeability and complex dielectric permittivity of the Co(2)Z ferrites substituted by Sr2+ ions. As Sr2+ ions increased, the real part of the magnetic permeability (μ′) hardly changed, whereas the imaginary part of the magnetic permeability (μ″) remained at a low value (~0.2 for all samples). According to the magnetic tangent tanδμ equation definition [23], as follows:
tanδμ = μ″/μ′,
an ultra-low order of magnitude of tanδμ (approximately 8 × 10−2) can be obtained over a long frequency band from 1 MHz to 1 GHz. tanδμ is a valuable factor for the antenna substrate materials. Herein, tanδμ is derived from three factors, namely the eddy current loss tangent tanδe, the hysteresis loss tangent tanδa and the remaining loss tangent tanδc; their relationship is as follows [8]:
tanδμ = tanδe + tanδa + tanδc
where tanδe generated from the electro-magnetic induction, causing cover fever and generating power dissipation, is closely correlated with the coercivity (Hc), and the Fe3O4 particles are scattered between the crystals. In addition, tanδe can be affected by existing pores and changing grain size. tanδa can be tailored by changing the hysteresis constant, which reduces the hysteresis loop and Hc. The remaining loss tangent tanδc mainly relies on the ideal microstructure of the materials with dense arrangement, uniform thickness, border crystal boundary and pores. In this research, the low magnetic loss is attributed to the low-temperature sintered technology, and the appropriate sintering aids Bi2O3. However, for complex dielectric permittivity, the actual part (ε′) increased from approximately 8 to 12 when the x value increased from 0.0 to 0.6. The dielectric constant increases with increased Sr2+ ions based on Koop’s phenomenological theory, in which the microstructure is regarded as a non-uniform intermediary of two layers, based on the Maxwell–Wagner type [36]. According to the theory, the dielectric construction of ferrites is composed of high- and low-conductivity grain grain boundaries. The high-conductivity grains are separated by grain boundaries, resulting in the localised build-up of charge carriers that increase interfacial polarisation. As a result, ε′ increases. As Sr3+ substitution content increases, the ability of the separation between the two layers is enhanced, and interfacial polarisation is further excited, resulting in an increase in ε′ [37]. In addition, reports showed a strong correlation between the conduction mechanism and the dielectric behaviour of the ferrites, starting with the supposition that the mechanism of the polarisation process in ferrites is similar to that of the conduction process. The electronic exchange in Fe2+⇔Fe3+ results in local displacement that determines the polarisation of the ferrites. More electronic exchange occurs in Fe2+⇔Fe3+ with the increase in Sr2+ ion content, resulting in higher local displacement that enhances dielectric properties [38]. Meanwhile, the imaginary part of the dielectric permittivity (ε″) was approximately 0.5 to 0.1 over the frequency of 1 MHz to 1 GHz. As a consequence, the dielectric loss tanδε was calculated (similar to that of tanδμ) to be approximately 0.007 in all samples. This value is also fairly low amongst ferrite ceramic materials. In general, tanδε is closely related to two factors: (i) the crystal grain boundaries and ferrite polycrystallinity and (ii) the microstructure including porosity and grain size. Their relationship can be expressed by the following equation [39]:
tanδε = (1 − P)tanδ0 + CPn
where tanδ0 is the dielectric loss of materials with completely dense microstructure, P is the porosity, and C is a dependent constant. In Equation (4), tanδε is mainly determined by P, which depends on relatively low porosity. In this work, high temperature enabled Bi2O3 to form molten BiFeO3 dielectric crystals that could fill the void between grains. Thus, the structures became denser, mainly causing low dielectric loss. In general, when tanδμ and tanδε are remarkably reduced, the proposed ferrite materials still exhibit excellent application prospects as antenna substrates.
The permeability and permittivity of all the samples in Figure 4 show that the actual parts of permeability and permittivity had similar values in the experimental scope. The miniaturisation factor (n, refractive index) and relative impedance (Z) of CO(2)Z ferrites with various Sr3+ ions substituted were calculated and are shown in Figure 5 and Table 1. The results show that n increases monotonically, and Z decreases with the increase in x value. However, Z decreases by much less than n increases. Finally, the sample substituted by Sr3+ ions with an x value of 0.6 exhibited the most ideal properties through comparison and consideration of various trade-offs.

4. Conclusions

In this study, superfluous Bi2O3 sintering aids were added to various Sr3+ ion-substituted CO(2)Z ferrites to achieve low-temperature sintering. The ferrites were made up of two phases and showed changing magnetic and dielectric properties, thereby approaching equal permeability and permittivity values. When the x value was 0.6, the materials had the topmost permittivity values (ε′ ≈ 12), whereas μ′ was barely changed. As a result, larger n and appropriate Z could be obtained. Meanwhile, low magnetic loss (tanδμ ≈ 0.08) and low dielectric loss (tanδε ≈ 0.007) indicate low power loss during operation. The results showed that the Sr3+ ions enhanced dielectric properties, indicating that CO(2)Z ferrite can be an excellent candidate for high-frequency antenna substrates.

Author Contributions

Conceptualization, J.W. and G.G.; methodology, K.L.; software, J.W.; validation, K.L. and G.G.; formal analysis, J.W.; investigation, J.W.; resources, K.L.; data curation, J.W.; writing—original draft preparation, J.W.; writing—review and editing, G.G.; visualization, K.L.; supervision, G.G.; project administration, K.L.; funding acquisition, G.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also form part of an ongoing study.

Conflicts of Interest

The authors declare no conflict of interest. We declare that we do not have any commercial or associative interest that represents a conflict of interest in connection with the work submitted.

References

  1. Saini, A.; Rana, K.; Thakur, A.; Thakur, P.; Mattei, J.L.; Queffelec, P. Low loss composite nano ferrite with matching permittivity and permeability in UHF band. Mater. Res. Bull. 2016, 76, 94–99. [Google Scholar] [CrossRef]
  2. Mattei, J.L.; Souriou, D.; Chevalier, A. Magnetic and dielectric properties in the UHF frequency band of half-dense Ni-Zn-Co ferrites ceramics with Fe-excess and Fe-deficiency. J. Magn. Magn. Mater. 2018, 447, 9–14. [Google Scholar] [CrossRef]
  3. Dar, M.A.; Varshney, D. Effect of d-block element Co2+ substitution on structural, Mössbauer and dielectric properties of spinel copper ferrites. J. Magn. Magn. Mater. 2017, 436, 101–112. [Google Scholar] [CrossRef]
  4. Peng, Y.; Wu, X.; Chen, Z.; Liu, W.; Wang, F.; Wang, X.; Feng, Z.; Chen, Y.; Harris, V.G. BiFeO3 tailored low loss M-type hexaferrite composites having equivalent permeability and permittivity for very high frequency applications. J. Alloys Compd. 2015, 630, 48–53. [Google Scholar] [CrossRef]
  5. Joulaei, M.; Hedayati, K.; Ghanbari, D. Investigation of magnetic, mechanical and flame retardant properties of polymeric nanocomposites: Green synthesis of MgFe2O4 by lime and orange extracts. Compos. Part B Eng. 2019, 176, 107345. [Google Scholar] [CrossRef]
  6. Reddy, M.P.; Shakoor, R.A.; Mohamed, A.M.A.; Gupta, M.; Huang, Q. Effect of sintering temperature on the structural and magnetic properties of MgFe2O4 ceramics prepared by spark plasma sintering. Ceram. Int. 2016, 42, 4221–4227. [Google Scholar] [CrossRef]
  7. Zu, Y.; Zhao, Y.; Xu, K.; Tong, Y.; Zhao, F. Preparation and comparison of catalytic performance for nano MgFe2O4, GO-loaded MgFe2O4 and GO-coated MgFe2O4 nanocomposites. Ceram. Int. 2016, 42, 18844–18850. [Google Scholar] [CrossRef]
  8. Peng, Y.; Wu, X.; Chen, Z.; Li, Q.; Yu, T.; Feng, Z.; Su, Z.; Chen, Y.; Harris, V.G. High frequency permeability and permittivity spectra of BiFeO3/(CoTi)-BaM ferrite composites. J. Appl. Phys. 2015, 117, 17A306. [Google Scholar] [CrossRef]
  9. Thakur, A.; Thakur, P.; Hsu, J.H. Novel magnetodielectric nanomaterials with matching permeability and permittivity for the very-high-frequency applications. Scr. Mater. 2011, 64, 205–208. [Google Scholar] [CrossRef]
  10. Yu, S.H.; Yoshimura, M. Ferrite/metal composites fabricated by soft solution processing. Adv. Funct. Mater. 2002, 12, 9–15. [Google Scholar] [CrossRef]
  11. Nasrin, S.; Khan, S.M.; Matin, M.A.; Khan, M.N.I.; Hossain, A.K.M.A.; Rahaman, M.D. Synthesis and deciphering the effects of sintering temperature on structural, elastic, dielectric, electric and magnetic properties of magnetic Ni0.25Cu0.13Zn0.62Fe2O4 ceramics. J. Mater. Sci. Mater. Electron. 2019, 30, 10722–10741. [Google Scholar] [CrossRef]
  12. Manhas, A.; Batoo, K.M.; Singh, M. Magnetic and Mössbauer investigations of soft Co2Z-type hexa nanoferrites. J. Alloys Compd. 2018, 767, 188–194. [Google Scholar] [CrossRef]
  13. Lee, W.; Hong, Y.K.; Park, J.; Larochelle, G.; Lee, J. Low-loss Z-type hexaferrite (Ba3Co2Fe24O41) for GHz antenna applications. J. Magn. Magn. Mater. 2016, 414, 194–197. [Google Scholar] [CrossRef]
  14. Xu, F.; Bai, Y.; Qiao, L.; Zhao, H.; Zhou, J. Realization of negative permittivity of Co2Z hexagonal ferrite and left-handed property of ferrite composite material. J. Phys. D Appl. Phys. 2009, 42, 025403. [Google Scholar] [CrossRef]
  15. Bae, S.; Hong, Y.K.; Lee, J.J.; Jalli, J.; Abo, G.S.; Lyle, A.; Seong, W.M.; Kum, J.S. Low loss Z-type barium ferrite (Co2 Z) for terrestrial digital multimedia broadcasting antenna application. J. Appl. Phys. 2009, 105, 07A515. [Google Scholar] [CrossRef]
  16. Chang, H.; Lee, H.B.; Chung, J.H.; Chun, S.H.; Shin, K.W.; Jeon, B.G.; Kim, K.H.; Prokeš, K.; Mat’aš, S. Commensurate transverse helical ordering in the room-temperature magnetoelectric Co2Z hexaferrite. Phys. B Condens. Matter. 2018, 551, 122–126. [Google Scholar] [CrossRef]
  17. Karimi-Maleh, H.; Karaman, C.; Karaman, O.; Karimi, F.; Vasseghian, Y.; Fu, L.; Baghayeri, M.; Rouhi, J.; Kumar, P.S.; Show, P.-L.; et al. Nanochemistry approach for the fabrication of Fe and N co-decorated biomass-derived activated carbon frameworks: A promising oxygen reduction reaction electrocatalyst in neutral media. J. Nanostruct. Chem. 2022, 12, 429–439. [Google Scholar] [CrossRef]
  18. Al Sharabati, M.; Abokwiek, R.; Al-Othman, A.; Tawalbeh, M.; Karaman, C.; Orooji, Y.; Karimi, F. Biodegradable polymers and their nano-composites for the removal of endocrine-disrupting chemicals (EDCs) from wastewater: A review. Environ. Res. 2021, 202, 111694. [Google Scholar] [CrossRef]
  19. Karimi-Maleh, H.; Khataee, A.; Karimi, F.; Baghayeri, M.; Fu, L.; Rouhi, J.; Karaman, C.; Karaman, O.; Boukherroub, R. A green and sensitive guanine-based DNA biosensor for idarubicin anticancer monitoring in biological samples: A simple and fast strategy for control of health quality in chemotherapy procedure confirmed by docking investigation. Chemosphere 2022, 291, 132928. [Google Scholar] [CrossRef]
  20. Yarahmadi, M.; Maleki-Ghaleh, H.; Mehr, M.E.; Dargahi, Z.; Rasouli, F.; Siadati, M.H. Synthesis and characterization of Sr-doped ZnO nanoparticles for photocatalytic applications. J. Alloys Compd. 2021, 853, 157000. [Google Scholar] [CrossRef]
  21. Hu, Z.Z.; Lu, J.J.; Chen, B.H.; Liu, X.Q.; Chen, X.M. Improved ferroelectric properties in hybrid improper ferroelectric Sr3−xBaxZr2O7. J. Alloys Compd. 2021, 866, 158705. [Google Scholar] [CrossRef]
  22. Gan, G.; Zhang, H.; Li, Q.; Li, J.; Li, M.; Xu, F.; Jing, Y. Bi2O3 enhances magnetic and dielectric properties of low temperature co-fired Ba(CoTi)1.20Fe9.6O19 ferrite composites in an oxygen atmosphere for applications in high frequency antennas. Mater. Res. Bull. 2018, 97, 37–41. [Google Scholar] [CrossRef]
  23. Gan, G.; Zhang, H.; Li, Q.; Li, J.; Huang, X.; Xie, F.; Xu, F.; Zhang, Q.; Li, M.; Liang, T.; et al. Low loss, enhanced magneto-dielectric properties of Bi2O3doped Mg-Cd ferrites for high frequency antennas. J. Alloys Compd. 2018, 735, 2634–2639. [Google Scholar] [CrossRef]
  24. Gan, G.; Zhang, D.; Li, J.; Wang, G.; Huang, X.; Rao, Y.; Yang, Y.; Wang, X.; Zhang, H.; Chen, R.T. Ga ions-tailored magnetic-dielectric properties of Mg–Cd composites for high-frequency, miniature and wideband antennas. Ceram. Int. 2020, 46, 8398–8404. [Google Scholar] [CrossRef]
  25. Gan, G.; Zhang, D.; Li, J.; Wang, G.; Huang, X.; Yang, Y.; Rao, Y.; Wang, X.; Zhang, H.; Chen, R.T. Low-loss Cd-substituted Mg ferrites with matching impedance for high-frequency-range antennas. J. Magnes. Alloy. 2021, 9, 1396–1405. [Google Scholar] [CrossRef]
  26. Li, Q.; Yan, S.; Wang, X.; Nie, Y.; Feng, Z.; Su, Z.; Chen, Y.; Harris, V.G. Dual-ion substitution induced high impedance of Co2Z hexaferrites for ultra-high frequency applications. Acta Mater. 2015, 98, 190–196. [Google Scholar] [CrossRef]
  27. Yang, P.; Qi, H.; Liu, Z.; Fu, X.; Peng, Z. Microstructure, magnetism, and high-frequency performance of polycrystalline Ni0.5Zn0.5Sm0.025HoxFe1.975−xO4 ferrites. J. Am. Ceram. Soc. 2019, 102, 7469–7479. [Google Scholar] [CrossRef]
  28. Li, Q.; Bao, S.; Liu, Y.; Li, Y.; Jing, Y.; Li, J. Influence of lightly Sm-substitution on crystal structure, magnetic and dielectric properties of BiFeO 3 ceramics. J. Alloys Compd. 2016, 682, 672–678. [Google Scholar] [CrossRef]
  29. Hao, P.; Zhao, Z.; Tian, J.; Sang, Y.; Yu, G.; Liu, H.; Chen, S.; Zhou, W. Bismuth titanate nanobelts through a low-temperature nanoscale solid-state reaction. Acta Mater. 2014, 62, 258–266. [Google Scholar] [CrossRef]
  30. Ghosh, M.P.; Mukherjee, S. Microstructural, magnetic, and hyperfine characterizations of Cu-doped cobalt ferrite nanoparticles. J. Am. Ceram. Soc. 2019, 102, 7509–7520. [Google Scholar] [CrossRef]
  31. Xu, F.; Liao, Y.; Zhang, D.; Zhou, T.; Li, J.; Gan, G.; Zhang, H. Synthesis of Highly Uniform and Compact Lithium Zinc Ferrite Ceramics via an Efficient Low Temperature Approach. Inorg. Chem. 2017, 56, 4512–4520. [Google Scholar] [CrossRef] [PubMed]
  32. Samajdar, R.; Scheurer, M.S.; Chatterjee, S.; Guo, H.; Xu, C.; Sachdev, S. Enhanced thermal Hall effect in the square-lattice Néel state. Nat. Phys. 2019, 15, 1290–1294. [Google Scholar] [CrossRef]
  33. Qiu, C.; Lu, T.; He, F.; Feng, S.; Fang, X.; Zuo, F.; Jiang, Q.; Deng, X.; Ye, J. Influences of gallium substitution on the phase stability, mechanical strength and cellular response of β-tricalcium phosphate bioceramics. Ceram. Int. 2020, 46, 16364–16371. [Google Scholar] [CrossRef]
  34. Trukhanov, A.V.; Trukhanov, S.V.; Kostishyn, V.G.; Panina, L.V.; Kazakevich, I.S.; Trukhanov, A.V.; Natarov, V.O.; Chitanov, D.N.; Turchenko, V.A.; Oliynyk, V.; et al. Microwave properties of the Ga-substituted BaFe12O19 hexaferrites. Mater. Res. Express 2017, 4, 076106. [Google Scholar] [CrossRef]
  35. Shinde, V.S.; Dahotre, S.G.; Singh, L.N. Synthesis and characterization of aluminium substituted calcium hexaferrite. Heliyon 2020, 6, e03186. [Google Scholar] [CrossRef]
  36. Naik, C.C.; Salker, A.V. Effect Cr3+ Ion Substitution on the Structural, Magnetic, and Dielectric Behavior of Co–Cu Ferrite. J. Supercond. Nov. Magn. 2019, 32, 3655–3669. [Google Scholar] [CrossRef]
  37. Sathishkumar, G.; Venkataraju, C.; Sivakumar, K. Magnetic and dielectric properties of cadmium substituted nickel cobalt nanoferrites. J. Mater. Sci. Mater. Electron. 2013, 24, 1057–1062. [Google Scholar] [CrossRef]
  38. Kumar, B.R.; Ravinder, D. Dielectric properties of Mn-Zn-Gd ferrites. Mater. Lett. 2002, 53, 437–440. [Google Scholar] [CrossRef]
  39. Krishnaveni, T.; Murthy, S.R.; Gao, F.; Lu, Q.; Komarneni, S. Microwave hydrothermal synthesis of nanosize Ta2O5 added Mg-Cu-Zn ferrites. J. Mater. Sci. 2006, 41, 1471–1474. [Google Scholar] [CrossRef]
Figure 1. SEM images of the materials with different Sr2+ ion substitution x values. (a) x = 0.0, (b) x = 0.2, (c) x = 0.4 and (d) x = 0.6.
Figure 1. SEM images of the materials with different Sr2+ ion substitution x values. (a) x = 0.0, (b) x = 0.2, (c) x = 0.4 and (d) x = 0.6.
Metals 12 01541 g001
Figure 2. XRD patterns of ferrites with different Sr2+ ion substitution x values.
Figure 2. XRD patterns of ferrites with different Sr2+ ion substitution x values.
Metals 12 01541 g002
Figure 3. Magnetic hysteresis loops. (a) Magnetic properties and (b) samples sintered with various Sr2+ ion substitution.
Figure 3. Magnetic hysteresis loops. (a) Magnetic properties and (b) samples sintered with various Sr2+ ion substitution.
Metals 12 01541 g003
Figure 4. Complex magnetic permeability. (a) Complex dielectric permittivity (b) and samples with various x values.
Figure 4. Complex magnetic permeability. (a) Complex dielectric permittivity (b) and samples with various x values.
Metals 12 01541 g004
Figure 5. Miniaturisation factor (n) and relative impedance (Z) of CO(2)Z ferrites with various x values.
Figure 5. Miniaturisation factor (n) and relative impedance (Z) of CO(2)Z ferrites with various x values.
Metals 12 01541 g005
Table 1. Miniaturisation factor (n, refractive index) and relative impedance (Z) of CO(2)Z ferrites corresponding to various x values.
Table 1. Miniaturisation factor (n, refractive index) and relative impedance (Z) of CO(2)Z ferrites corresponding to various x values.
x Value0.000.200.400.60
n value4.5555.2
Z value0.560.500.500.48
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, J.; Li, K.; Gan, G. Sr2+ Ion Substitution Enhanced Dielectric Properties of Co(2)Z Ferrites for VHF Antenna Substrate. Metals 2022, 12, 1541. https://doi.org/10.3390/met12091541

AMA Style

Wang J, Li K, Gan G. Sr2+ Ion Substitution Enhanced Dielectric Properties of Co(2)Z Ferrites for VHF Antenna Substrate. Metals. 2022; 12(9):1541. https://doi.org/10.3390/met12091541

Chicago/Turabian Style

Wang, Ji, Kunlong Li, and Gongwen Gan. 2022. "Sr2+ Ion Substitution Enhanced Dielectric Properties of Co(2)Z Ferrites for VHF Antenna Substrate" Metals 12, no. 9: 1541. https://doi.org/10.3390/met12091541

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