Next Article in Journal
Influences of Personal Driving Styles and Experienced System Characteristics on Driving Style Preferences in Automated Driving
Previous Article in Journal
The Antibacterial Effect of the Films Coated with the Layers Based on Uncaria tomentosa and Formitopsis betulina Extracts and ZnO Nanoparticles and Their Influence on the Secondary Shelf-Life of Sliced Cooked Ham
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetoelectric Composites: Engineering for Tunable Filters and Energy Harvesting Applications

1
Faculty of Science and Technology, Institute of Materials Engineering, University of Silesia in Katowice, 75 Pułku Piechoty 1a, 41-500 Chorzów, Poland
2
Empa—Swiss Federal Laboratories for Materials Science and Technology, Department of Functional Materials, Überlandstrasse 129, 8600 Dübendorf, Switzerland
3
Materials Processing Institute, Eston Road, Middlesbrough TS6 6US, UK
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(15), 8854; https://doi.org/10.3390/app13158854
Submission received: 4 July 2023 / Revised: 25 July 2023 / Accepted: 28 July 2023 / Published: 31 July 2023

Abstract

:
Multiferroic ceramic composites have been engineered to incorporate multiple desired physical properties within a single ceramic component. The objective of this study was to create such composites through pressure less sintering ferroelectric-doped PZT and nickel–zinc ferrite at a temperature of 1250 °C. The growth of ferrite grains was found to be influenced by the concentration of the ferroelectric PZT phase. Consequently, an increase in the ferrite content decreased the average particle size of nickel–zinc ferrite by a factor of 1.8. After impedance spectroscopy, the multiferroic ceramic composites can be categorized into two groups: those with low ferrite content (<20%) and those with a high ferrite content (>20%). Composites with a high ferrite content are suitable for dual-band filters or shield applications. The impedance spectroscopy analysis revealed that the resonance frequency can be shifted to higher frequency ranges. Therefore, it was demonstrated that modifying the composition of the multiferroic composite allows for tailoring the impedance behavior to shield living and working spaces against such radiation to meet the demands of the 21st century.

1. Introduction

In a constantly evolving world of modern techniques and microelectronics, development of materials with wide, functional properties is imperative [1,2,3,4]. The group of materials with interesting properties includes, among others, ferroics, multiferroic, piezoelectric, pyroelectric, multicomponent solid solutions, and multiferroic ceramic composites [5,6,7,8,9]. Previously, multiferroic materials have been widely investigated because of their extensive applications in multifunctional applications such as non-volatile memories that can be electrically written and magnetically read [10], spin-orbit torque field-effect transistors (SOTFET) that can operate as a non-volatile memory and logic device [11] or electrically tunable microwave devices [12]. Multiferroic composites’ prospective applications (e.g., ferroelectric–ferromagnetic composites) mainly depend on the coupling of magnetic and electrical substructures [13,14,15]. Modern ceramic filters, including multiferroic ceramic composites, are designed to acquire multiple (appropriate) physical properties in a single ceramic element, which is conducive to maintaining the miniaturization of microelectronic elements [16].
Additionally, most wearable electronics use a lithium-ion battery, which causes severe problems with sustainability. Actual solutions are multiferroic energy harvesters that can convert vibration and biomechanical energy to power personal electronics in an environmentally friendly way. However, to maximize its effectiveness, it must also be tuned to the frequency of the runner [17].
For many decades, the main material used in the construction of functional materials is a solid solution of (x)PbTiO3-(1-x)PbZrO3 (PZT) due to its excellent dielectric, piezoelectric, pyroelectric, and non-linear optical properties [18,19,20,21,22]. This type of suitable property is obtained via appropriate doping of the basic PZT composition and the use of proper methods of a technological process [23,24,25,26]. Especially, chemical compositions of PZT, close to the morphotropic boundary (mixture of a rhombohedral and a tetragonal phase), exhibit optimal and excellent physical properties that are suitable for modern microelectronic applications, e.g., for actuators, piezo-transducers, sensors, accelerometers, computer memories, and MEMS devices [27,28,29]. Interesting and new functional properties of multiferroic materials can be obtained as ceramic composites by combining the piezoelectric and magnetic components [30]. This creates new possibilities for their versatile use [31,32,33,34]. Magnetic ferrites (e.g., nickel–zinc, manganese–zinc, or manganese–chromium–zinc ferrites) are used for this type of connection, and the magnetic properties of the ferrites depend on their chemical composition [35,36].
In this work, a study set of six multiferroic composites with two main phases, the ferroelectric phase P (Bi, Nb, and Mn-doped PZT material) and the magnetic one, i.e., nickel–zinc ferrite F, were performed. The ferroelectric P material was selected based on its high dielectric and ferroelectric properties. In contrast, in the choice of magnetic F component, the ferrite’s relatively high resistance and appropriately high magnetic properties were decisive. The composite samples had the following content of their components, P/F: 90/10 (P90-F10), 85/15 (P85-F15), 80/20 (P80-F20), 60/40 (P60-F40), 40/60 (P40-F60), and 20/80 (P20-F80). A previous paper [37] investigated X-ray diffraction patterns, microstructural, ferroelectric, dielectric, magnetic properties, and DC electrical conductivity of the P-F composite materials. Thanks to these properties, these materials can be used in modern microelectronic application. The research presented in the work [37] was conducted to search for a composition with the best properties for new types of memory and electromagnetic converters. The research in the present article focuses on the effect of nickel–zinc ferrite content on the microstructure using the backscattered electron detector (BSE) and by blocking the electromagnetic field (EMF) at low and high frequencies; at the same time impedance spectroscopy of the P-F multiferroic composite materials were systematically investigated.
The idea of engineering magnetoelectric ferrite–ferroelectric composites for application as tunable electromagnetic filters were first proposed by Ciomaga et al. and finally confirmed by Atif et al.; the results of these studies were in agreement with the earlier numerical simulations also provided by Cristina E Ciomaga et al. [38,39]. It is worth noting that recent achievements in the field of filtering ionizing radiation also successfully use multiferroic solid-state composites for this purpose [40]. Consequently, the present research aims to resolve the complex problems with structural compatibility of the ferrite–ferroelectric components in order to reduce the internal stresses and porosity at the interfaces between participant phases to ensure magnetoelectric coupling between the magnetic and ferroelectric phases.

2. Materials and Methods

Multiferroic ceramic composites, P-F, based on ferroelectric and magnetic materials were fabricated using a conventional solid-state reaction method and different ratios of ferroelectric and ferromagnetic components. The individual components of the composite were selected based on their high dielectric and ferroelectric properties (P) and good magnetic properties (F). The ferroelectric component was doped using a PZT solid solution with the chemical formula Pb(Zr0.51Ti0.49)O3 + 0.2%at. Bi2O3 + 0.03%at. Nb2O5 + 0.06%at. MnO2 (P), while the magnetic component was a ferrite material with the chemical composition Ni0.5Zn0.5Fe2O4 (F). The input ingredients to obtain the two components of the composite were high-purity simple and complex compounds. The detailed course of the technological process of the ferroelectric PZT-type material and ferrite material is given in a previous work [37].
The synthesized composite components, i.e., P and F powders, were weighed in appropriate proportions (P/F ratios) and then mixed in a Fritsch Pulverisette-6 (Idar-Oberstein, Germany) planetary ball mill (wet milling in ethyl alcohol for 15 h). The powders mixtures were synthesized via calcination under the condition of 950 °C/2 h, while sintering of the multiferroic composite samples was carried out via the free sintering (pressureless) method (1250 °C/2 h). For electrical research, the surfaces of samples were polished, and silver electrodes were applied to them. The analysis of the results was performed on a series of multiferroic composite samples (P-F) with the following P/F percentages: 20/80 (P20-F80), 40/60 (P40-F60), 60/40 (P60-F40), 80/20 (P80-F20), 85/15 (P85-F15), and 90/10 (P90-F10).
The observation of the surface morphology and the chemical composition tests of the multiferroic composite samples were recorded on a scanning electron microscope in combination with an energy dispersive spectrometer SEM-EDS (JEOL JSM-7100F TTL LV from Jeol Ltd., Tokyo, Japan). Two image capture techniques have been used: (i) the standard SB method, i.e., combining signals with the secondary and backscattered electron detectors, and (ii) the BSE method, i.e., signal from backscattered electron detector. The average grain size has been estimated using the ImageJ program. The X-ray diffraction patterns were tested at room temperature using the X’Pert Pro diffractometer (PANalytical, Eindhoven, the Netherlands) in the 2θ range 10–60°. Frequency measurements of the electric parameters were conducted on the 1260 A Frequency Response Analyzer, which is the cornerstone of FRA technology (AMETEK Scientific Instruments Berwyn, PA, USA) in a frequency range from 0.1 Hz to 10 MHz and at a temperature range from 30 to 400 °C.

3. Results

Figure 1 depicts SEM images of microstructures of the P-F multiferroic composite samples created using two image capture techniques, i.e., SB (Figure 1a–f) and BSE (Figure 1a′–f′), for the same analysis area of the surface composite samples. BSE enables convenient visualization of the distribution of individual components on the surface of the composite sample, i.e., areas with magnetic (dark regions) and ferroelectric (bright regions) phases. By filtering out low-energy secondary electrons, the BSE analysis reveals bright areas with a predominance of grains with a higher atomic number. In comparison, the dark grains indicate areas with a predominance of elements with a lower atomic number. The assignment of grains to the appropriate phase of the composite was confirmed via spot EDS analysis. The presented photos show how the appearance of the microstructure of the analyzed series of multiferroic composite samples with the change in the ferrite (ferroelectric) phase evolves. The PZT-type material (ferroelectric phase) shows a high degree of solidification; the grains are fine and firmly solidified with each other, with precise edges. On the other hand, magnetic grains are usually larger and are also correctly crystallized with clear grain boundaries. SEM images show that ferroelectric and magnetic grains are evenly distributed in the composite samples’ microstructure of the cross-section. However, not all magnetoelectric materials and composites can obtain a homogeneous microstructure in the technological process, e.g., [41,42,43]. In the case of compositions with a dominant ferroelectric phase (constituting the matrix of the multiferroic composite), fine grains of the ferroelectric component surround the larger magnetic grains (P90-F10 sample). In many cases, the magnetic grains tend to aggregate into larger clusters of the magnetic fraction. When the amount of a magnetic phase in composite composition increases, the situation reverses. For the P20-F80 composition, numerous magnetic grains surrounding small amounts of the ferroelectric component distributed randomly.
Figure 2 shows the grain size distribution of the P-F multiferroic composite materials. In general, magnetic materials have lower sintering temperatures than PZT materials; hence, the growth of magnetic grains starts at a lower temperature. In composite compositions with a significant predominance of the ferroelectric component, the magnetic grains have more space suitable for their expansion, which results in a larger average grain size. Ferroelectric grains grow similarly in these compositions. In the case of composite compositions with a dominant magnetic phase, the overwhelming number of magnetic grains hinders their excessive growth and promotes the improvement in grain size homogeneity. However, ferroelectric grains in these compositions have difficult growth conditions, resulting in low average grain size.
As it is known, the physical (dielectric) properties of multiferroic materials depend on their microstructure, including homogeneity, grain size, and density. The microstructure, in turn, is influenced by an adequately conducted technological process of obtaining multiferroic materials (including mixing powders, the appropriate selection of the method, and conditions of synthesis and sintering). In general, the homogeneity of the microstructure of ceramic materials has a positive effect on their final properties, and in the case of dielectric properties, on the degree of diffusion of the phase transition (ferroelectric/paraelectric). Ceramic materials with high uniformity of microstructure show a sharp phase transition. Grain size also affects the dielectric properties of ceramic materials. Materials with smaller grains show a stronger effect of mechanical stress and an external effect of grain boundaries, which result in increased blurring of the phase transition, lower Curie temperature, and lower permittivity [44,45]. The dielectric constant of the materials is linearly proportional to the average value of grain size and inversely proportional to the grain boundary thickness (as grain size increases, permittivity increases) [46].
The EDS analyses of the P-F multiferroic composites are presented in Figure 3a–f (according to [37]), and the given values are the average values from five measurement areas. The calculations of the theoretical (t) and experimental (ex) results of oxides in the composite composition are summarized in Table 1. The EDS investigation confirmed no presence of foreign elements. In the case of compositions with a predominant ferroelectric phase content, high compliance with the theoretical calculations for individual composite compositions is observed. In the case of compositions with a predominant content of the magnetic phase, the analysis showed an excess content of the magnetic ferrite phase.
The electron probe microanalyses (EPMA) represent the distribution of the composite elements on the surface of the multiferroic ceramic sample. Figure 3g shows a representative EPMA analysis for the 90P-10F composite sample. The maps of individual elements show an increase in signal detection intensity in areas richer in this element and vice versa. As a result, the obtained map provides a complete picture of the distribution of the magnetic and ferroelectric components on the surface of the composite sample. In the case of the elements with the lowest content, the measurement is the least accurate. The EPMA complies with SEM imaging using the BSE technique (Figure 1f′), confirming the effectiveness of both methods of microstructure analysis. The full EPMA results for multiferroic P-F composites are presented in [37].
The X-ray diffraction patterns of the multiferroic composites tested at room temperature are presented in Figure 4. The analysis showed the presence of firm peaks in the ferroelectric phase (coexistence of two tetragonal and rhombohedral phases) and the magnetic phase (ferrite). In the case of the P ferroelectric phase, the best match was obtained for the card no. 00-33-0784 (for the tetragonal crystal system with space group P4mm) and the card no. 01-073-2022 (for the rhombohedral crystal system with space group R3m). On the other hand, the best match for ferrite materials (F) was obtained from the JCPDS data 08–0234 (for the spinel phase with space group Fd-3m). Depending on the change in individual components, a change in the intensity of individual reflections is observed.
The dielectric measurements presented in previous work [37] showed that in the tested material, there is a clear tendency for the electrical loss to increase with an increase in the ferromagnetic phase content. However, such a tendency is not reflected in the subsequent permittivity spectra. To explain the reasons for the above dielectric phenomena, impedance spectroscopy (IS) was implemented to study the local maxima behavior of P-F multiferroic composites. IS test is a very effective tool for thoroughly analyzing the relaxations in many materials [47].
Figure 5 shows the complex impedance plots of Z″ versus Z′ (Cole-Cole plots) from the temperature range 40–400 °C of the investigated ceramics with 10, 15, 20, 40, 60, and 80% ferrite content. For lower ferrite content samples, i.e., 10, 15, and 20% (P90-F10, P85-F15, P80-F20) samples), we had to numerically separate the not apparent semicircles on the plots at each temperature because the two semicircular arcs were very close due to similar grain and grain boundary impedance levels. We assigned the separated high-frequency arc to the grain interior, and the low-frequency arc to the grain boundary [48]. Completely different spectra were recorded for the samples with a higher ferrite content. Mainly two well separated semicircular were recorded for the composites with 80, 60, and 40% ferrite content (P20-F80, P40-F60, and P60-F40), presented in Figure 5a–c, respectively. Such a shape indicates that the ferrite phase dramatically decreases the impedance of the grain interior, which is consistent with the dielectric measurements mentioned above [49,50]. Generally, for all the compositions, the diameter of semicircles lowered significantly with the increase in measuring temperature, which illustrated a strong temperature dependence of the conductivity mechanism. In our opinion, this effect indicates that the mobility of the space charge became easier, and the accumulated charge carriers in the vicinity of the grain interior had sufficient energy to pass through the ferrite phase, which led to an enhanced conductivity with the reduction in impedance that was also presented in [51,52].
In correlation to the presented complete temperature range characteristics, the more detailed impedance spectra recorded at the temperature range from 200 to 300 °C of the P-F multiferroic composites are presented in Figure 6. Similarly, two semicircular arcs are very distinctly visible in the Nyquist plot of high ferrite content samples in low and high frequencies, indicating the presence of two parallel relaxation processes (Figure 6a–c).
Although for almost all compositions there is a weaker grain boundary effect reflected in the shape of the smaller semicircle at low frequency, for the P60-F40 sample, the dominant is the grain effect (Figure 6c) [53]. A deviation from the double semicircle form of the P-F composite samples has been recorded with decreasing ferrite content (Figure 6d–f).
In order to calculate the activation energies of these relaxations, the additional charts are constructed and presented in Figure 7 that showed the frequency dependence of the normalized imaginary part of the impedance calculated as a Z″ to Zmax ratio in the same temperature range as in Figure 6. The most exciting thing that is visible is the apparent frequency shift to a higher frequency range of the grain boundaries’ peaks with an increasing amount of ferrite phase in the investigated P-F composite material. For the sample with only 10% of ferrite content (P90-F10), the frequency range is from 100 Hz to 3 kHz, whereas, for example, for the sample with 40% of ferrite content analogical range is from 1 kHz to 12 kHz. Consequently, we can draw significant conclusions from the application’s point of view because we can tune the resonant frequency with composition. Hence, in the case of antennas or band filters, we can fabricate appropriate filters for a specific frequency band through material engineering. Only for the 10, 15, and 20% ferrite content samples (P90-F10, P85-F15, and P80-F20, respectively), single Debye-like peaks are visible, whereas a secondary peak has appeared for a higher ferrite content. From these peaks, the relaxation time can be obtained as the inverse of fmax:
τ = 1 2 π f m a x
Consequently, from the plots of ln(τ) vs. 1/T, that obey the Arrhenius equation, the ER value was calculated for the investigated P-F composite materials and listed in Table 2. Considering the obtained ER values for these composites, it seems that charge movement involves only released electron conductivity because for oxygen vacancies VO the activation energy for VO is much higher, namely in the range of 0.7–1.1 eV [54].
The Arrhenius law can express the activation energy of conductivity [55],
σ A C = σ 0 exp ( E a / k B T )
where σAC, σ0, Ea, T, and kB, are conductivity, activation energy, temperature, and Boltzmann constant, respectively. The recorded data were fitted with the above equation, and the resulting values of activation energies are shown in Figure 8. For the pure PZT material, the value of activation energy was taken from theoretical calculations presented in the work of Betul, Akkopru-Akgun, and Susan Trolier-McKinstry [56]. Figure 8a shows the activation energy values for the studied P-F multiferroic samples and grain and grain boundary conductivity contributions (other for the grains and grain boundaries). In the 20% of ferrite phase content sample (P80-F20 sample), there is a significant difference in the level of activation energies which means that the ferrite conductive phase is mainly located in grain boundaries. Consequently, the grain contribution to the complex AC response is lower than those of the results for the P-F composite samples with higher than 20% of ferrite content (i.e., P60-F40, P40-F60, and P20-F80 samples), where the grain and grain boundary contributions are almost equal. Figure 8b demonstrates that our multiferroic tunable bandpass filters are very precise and efficient, which could have many applications in energy harvesting and tunable bandpass filters.

4. Conclusions

Ferroelectric–ferromagnetic PZT-NZF ceramic composites (P-F) for tunable dual-band energy harvesters and electromagnetic as well as ionizing radiation shielding were successfully synthesized using the conventional mixed oxide method, and the correlation of microstructural structural properties was evaluated and presented.
The SEM of obtained ferroelectric–ferromagnetic ceramic composite microstructure revealed perfect grain arrangement and the XRD examination of the crystallographic structure registered profiles of both ferroelectric and magnetic phases.
The most important results are the shifts in the impedance spectra of the maximum absorption peaks of electromagnetic waves, which served as a qualitative experimental proof of the possibility of using engineering magnetoelectric ferrite–ferroelectric composites as tunable electromagnetic filters.
In particular, the results obtained for the P40-F60 multiferroic composite offer a new field of application for new piezoelectric–ferrite structures with double blocking band properties. An additional attractive property discovered is the compositional tunability, an apparent frequency shift to a higher frequency range with an increase in the amount of ferrite phase in the tested composite material over a wide frequency range from 100 mHz to 1 MHz.
In summary, we have obtained a P-F network that enables dual-band characteristics and have established a method that modifies these properties with two impedance peaks for dual-band filters, EMF and ionizing radiation shields or energy harvesters.

Author Contributions

Conceptualization, L.K.; methodology, D.B.; validation and formal analysis, F.C. and T.S. All authors have read and agreed to the published version of the manuscript.

Funding

The present paper in part was financed by the Polish Ministry of Education and Science within statutory activity and in part by “Student Science Clubs Create Innovations” program [grant number SKN/SP/535865/2022].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The present paper in part was financed by the Polish Ministry of Education and Science within statutory activity and in part by “Student Science Clubs Create Innovations” program (grant number SKN/SP/535865/2022).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Henrique, M.B., Jr.; Magalhães, A.L.; Bastos, A.M.; Jefferson, M.; Sales, A.; Gouveia, D.X.; Kamshilin, A.A.; da Ferreira, A.C.; de Guimarães, G.F.; Sombra, A.S.B.; et al. Piezoelectric ceramic sensor (PZT) applied to electric current measurements. Microsyst. Technol. 2019, 25, 705–710. [Google Scholar]
  2. Hu, Z.; Smith, R.C.; Ernstberger, J. The homogenized energy model for characterizing polarization and strains in hysteretic ferroelectric materials: Implementation algorithms and data-driven parameter estimation techniques. J. Intel. Mat. Syst. 2012, 16, 1869–1894. [Google Scholar] [CrossRef]
  3. Aktas, O.; Kangama, M.; Linyu, G.; Catalan, G.; Ding, X.; Zunger, A.; Salje, E.K.H. Piezoelectricity in nominally centrosymmetric phases. Phys. Rev. Res. 2021, 3, 043221. [Google Scholar] [CrossRef]
  4. Chen, D.; Wang, L.; Luo, X.; Fei, C.; Li, D.; Shan, G.; Yang, Y. Recent development and perspectives of optimization design methods for piezoelectric ultrasonic transducers. Micromachines 2021, 12, 779. [Google Scholar] [CrossRef]
  5. Meng, Y.; Chen, G.; Huang, M. Piezoelectric materials: Properties, advancements, and design strategies for high-temperature applications. Nanomaterials 2022, 12, 1171. [Google Scholar] [CrossRef]
  6. Dipti, D.; Juneja, J.K.; Singh, S.; Raina, K.K.; Prakash, C. Enhancement in magnetoelectric coupling in PZT based composites. Ceram. Int. 2015, 41, 6108–6112. [Google Scholar] [CrossRef]
  7. Spaldin, N.A. Multiferroics: Past, present, and future. MRS Bull. 2017, 42, 385–389. [Google Scholar] [CrossRef] [Green Version]
  8. Bartkowska, J.A.; Ilczuk, J. Microscopic origin of electric and magnetic ordering in BiFeO3. Int. J. Thermophys. 2010, 31, 1–7. [Google Scholar] [CrossRef]
  9. Bochenek, D.; Skulski, R.; Wawrzała, P.; Brzezińska, D. Dielectric properties of Pb0.75Ba0.25(Zr0.65Ti0.35)1-zSnzO3 ceramics. Ferroelectrics 2011, 418, 82–87. [Google Scholar] [CrossRef]
  10. Scott, J.F. Multiferroic memories. Nat. Mater. 2007, 6, 256–257. [Google Scholar] [CrossRef]
  11. Afuye, O.; Li, X.; Guo, F.; Jena, D.; Ralph, D.C.; Molnar, A.; Xing, H.G.; Apsel, A. Modeling and circuit design of associative memories with spin orbit torque FETs. IEEE J. Explor. Solid-State Comput. Dev. Circ. 2019, 5, 197–205. [Google Scholar] [CrossRef]
  12. Ustinov, A.B.; Kalinikos, B.A. Multiferroic periodic structures based on magnetic crystals for electronically tunable microwave devices. Tech. Phys. Lett. 2014, 7, 568–570. [Google Scholar] [CrossRef]
  13. Kreisel, J.; Kenzelmann, M. Multiferroics the challenge of coupling magnetism and ferroelectricity. Europhys. News 2009, 40, 17–20. [Google Scholar] [CrossRef]
  14. Fiebig, M. Revival of the magnetoelectric effect. J. Phys. D Appl. Phys. 2005, 38, R123–R152. [Google Scholar] [CrossRef]
  15. Das, J.; Li, M.; Kalarickal, S.S.; Altmannshofer, S.; Buchanan, K.S.; Li, J.F.; Viehland, D. Control of magnetic and electric responses with electric and magnetic fields in magnetoelectric heterostructures. Appl. Phys. Lett. 2010, 96, 222508. [Google Scholar] [CrossRef] [Green Version]
  16. Yang, G.M.; Lou, J.; Wu, J.; Liu, M.; Wen, G.; Jin, Y.; Sun, N.X. Dual H-and E-field tunable multiferroic bandpass filters with yttrium iron garnet film. In Proceedings of the 2011 IEEE MTT-S International Microwave Symposium, Baltimore, MD, USA, 5–10 June 2011; pp. 1–4. [Google Scholar]
  17. Wu, H.; Tatarenko, A.; Bichurin, M.I.; Wang, Y. A multiferroic module for biomechanical energy harvesting. Nano Energy 2021, 83, 105777. [Google Scholar] [CrossRef]
  18. Xiang, P.-H.; Dong, X.-L.; Chen, H.; Zhang, Z.; Guo, J.-K. Mechanical and electrical properties of small amount of oxides reinforced PZT ceramics. Ceram. Int. 2003, 29, 499–503. [Google Scholar] [CrossRef]
  19. Boucher, E.; Guiffard, B.; Lebrun, L.; Guyomar, D. Effects of Zr/Ti ratio on structural, dielectric and piezoelectric properties of Mn- and (Mn, F)-doped lead zirconate titanate ceramics. Ceram. Int. 2006, 32, 479–485. [Google Scholar] [CrossRef]
  20. Scott, J.F. Applications of magnetoelectrics. J. Mater. Chem. 2012, 22, 4567–4574. [Google Scholar] [CrossRef]
  21. Bochenek, D.; Niemiec, P.; Skulski, R.; Adamczyk, M.; Brzezińska, D. Electrophysical properties of the multicomponent PBZT-type ceramics doped by Sn4+. J. Electroceram. 2019, 42, 17–30. [Google Scholar] [CrossRef] [Green Version]
  22. Brzezińska, D.; Skulski, R.; Bochenek, D.; Niemiec, P. The properties of (1-x)(0.5PZT-0.5PFW)-xPFN ceramics. Integr. Ferroelectr. 2016, 173, 104–112. [Google Scholar] [CrossRef]
  23. Venkata Ramana, M.; Roopas Kiranc, S.; Ramamanohar Reddy, N.; Siva Kumar, K.V.; Murthy, V.R.K.; Murty, B.S. Investigation and characterization of Pb(Zr0.52Ti0.48)O3 nanocrystalline ferroelectric ceramics: By conventional and microwave sintering methods. Mater. Chem. Phys. 2011, 126, 295–300. [Google Scholar] [CrossRef]
  24. Maiwa, H.; Kimura, O.; Shoji, K.; Ochiai, H. Low temperature sintering of PZT ceramics without additives via an ordinary ceramic route. J. Eur. Ceram. Soc. 2005, 25, 2383–2385. [Google Scholar] [CrossRef]
  25. Xu, Z.J.; Chu, R.Q.; Li, G.R.; Shao, X.; Yin, Q.R. Preparation of PZT powders and ceramics via a hybrid method of sol–gel and ultrasonic atomization. Mater. Sci. Eng. B 2005, 117, 113–118. [Google Scholar] [CrossRef]
  26. Zhang, L.; Jiang, S.; Fan, B.; Zhang, G. Enhanced energy storage performance in (Pb0.858Ba0.1La0.02Y0.008) (Zr0.65Sn0.3Ti0.05)O3–(Pb0.97La0.02) (Zr0.9Sn0.05Ti0.05)O3 anti–ferroelectric composite ceramics by Spark Plasma Sintering. J. Alloys Compd. 2015, 622, 162–165. [Google Scholar] [CrossRef]
  27. Prasatkhetragarn, A.; Yimnirun, R. Phase formation, electrical properties and morphotropic phase boundary of 0.95Pb(ZrxTi1-x)O3–0.05Pb(Mn1/3Nb2/3)O3 ceramics. Ceram. Int. 2013, 39, S91–S95. [Google Scholar] [CrossRef]
  28. Li, J.; Sun, Q. Effects of Cr2O3 doping on the electrical properties and the temperature stabilities of PZT binary piezoelectric ceramics. Rare Met. 2008, 27, 362–366. [Google Scholar] [CrossRef]
  29. Shung, K.K.; Cannata, J.M.; Zhou, Q.F. Piezoelectric materials for high frequency medical imaging applications: A review. J. Electroceram. 2007, 19, 141–147. [Google Scholar] [CrossRef]
  30. Venkata Ramana, M.; Ramamanohar Reddy, N.; Siva Kumar, K.V.; Murthy, V.R.K.; Murty, B.S. Magneto-electric effect in multiferroic Ni0.93Co0.02Mn0.05Fe1.95O4−/PbZr0.52Ti0.48O3 particulate composites: Dielectric, piezoelectric properties. Mod. Phys. Lett. B 2011, 25, 345–358. [Google Scholar] [CrossRef]
  31. Zhang, H.; Mak, C.-L. Impedance spectroscopic characterization of fine-grained magnetoelectric Pb(Zr0.53Ti0.47)O3–(Ni0.5Zn0.5)Fe2O4 ceramic composites. J. Alloys Compd. 2012, 513, 165–171. [Google Scholar] [CrossRef]
  32. Němec, P.; Fiebig, M.; Kampfrath, T.; Kimel, A.V. Antiferromagnetic opto-spintronics. Nat. Phys. 2018, 14, 229–241. [Google Scholar] [CrossRef]
  33. Hrib, L.M.; Caltun, O.F. Effects of the chemical composition of the magnetostrictive phase on the dielectric and magnetoelectric properties of cobalt ferrite–barium titanate composites. J. Alloys Compd. 2011, 509, 6644–6648. [Google Scholar] [CrossRef]
  34. Bochenek, D.; Zachariasz, R.; Niemiec, P.; Bartkowska, J.; Brzezińska, D. Ferroelectromagnetic solid solutions on the base piezoelectric ceramic materials for components of micromechatronics. Mech. Syst. Signal Process. 2016, 78, 84–90. [Google Scholar] [CrossRef]
  35. Penchal Reddy, M.; Madhuri, W.; Venkata Ramana, M.; Ramamanohar Reddy, N.; Siva Kumar, K.V.; Murthy, V.R.K.; Siva Kumar, K.; Ramakrishna Reddy, R. Effect of sintering temperature on structural and magnetic properties of NiCuZn and MgCuZn ferrites. J. Magn. Magn. Mater. 2010, 322, 2819–2825. [Google Scholar] [CrossRef]
  36. Penchal Reddy, M.; Madhuri, W.; Ramamanohar Reddy, N.; Siva Kumar, K.V.; Murthy, V.R.K.; Ramakrishna Reddy, R. Magnetic properties of Ni-Zn ferrites prepared by microwave sintering method. J. Electroceram. 2012, 28, 1–9. [Google Scholar] [CrossRef]
  37. Bochenek, D.; Niemiec, P.; Chrobak, A. Effect of chemical composition on magnetic and electrical properties of ferroelectromagnetic ceramic composites. Materials 2021, 14, 2488. [Google Scholar] [CrossRef]
  38. Ciomaga, C.E.; Avadanei, O.G.; Dumitru, I.; Airimioaei, M.; Tascu, S.; Tufescu, F.; Mitoseriu, L. Engineering magnetoelectric composites towards application as tunable microwave filters. J. Phys. D Appl. Phys. 2016, 49, 125002. [Google Scholar] [CrossRef]
  39. Atif, M.; Ahmed, S.; Nadeem, M.; Nasir Khan, M. Complex dielectric and impedance analysis in a relaxor type ferroelec-tric/ferrimagnetic magnetoelectric (0.5)PbZr0.52Ti0.48O3+(0.5)CoFe2O4 composite. J. Alloys Compd. 2018, 735, 880–889. [Google Scholar] [CrossRef]
  40. Hannachi, E.; Sayyed, M.I.; Slimani, Y.; Almessiere, M.A.; Baykal, A.; Elsafi, M. Synthesis, characterization, and performance as-sessment of new composite ceramics towards radiation shielding applications. J. Alloys Compd. 2022, 899, 163173. [Google Scholar] [CrossRef]
  41. Wu, H.; Xu, R.; Zhou, C.; Xing, S.; Zeng, Z.; Ao, H.; Li, W.; Qin, X.; Gao, R. Effect of core size on the magnetoelectric properties of Cu0.8Co0.2Fe2O4@Ba0.8Sr0.2TiO3 ceramics. J. Phys. Chem. Solids 2022, 160, 110314. [Google Scholar] [CrossRef]
  42. Gao, R.; Zhang, Q.; Xu, Z.; Wang, Z.; Chen, G.; Deng, X.; Fu, C.; Cai, W. A comparative study on the structural, dielectric and multiferroic properties of Co0.6Cu0.3Zn0.1Fe2O4/Ba0.9Sr0.1Zr0.1Ti0.9O3 composite ceramics. Compos. Part B-Eng. 2019, 166, 204–212. [Google Scholar] [CrossRef]
  43. Mane, S.M.; Nimbalkar, A.R.; Kim, H.; Kulkarni, S.B.; Tayade, N.T.; Thombare, J.V.; Dhasade, S.S.; Shin, J.C. Magnetoelectric and magnetodielectric coupling in partially Ni-doped CoFe2O4 and 0.15(Ba0.7Ca0.3TiO3)-0.85(BaZr0.2Ti0.8O3) composites prepared via clean microwave sintering. J. Alloys Compd. 2020, 849, 156599. [Google Scholar] [CrossRef]
  44. Mao, C.; Yan, S.; Cao, S.; Yao, C.; Cao, F.; Wang, G.; Dong, X.; Hu, X.; Yang, C. Effect of grain size on phase transition, dielectric and pyroelectric propertiesof BST ceramics. J. Eur. Ceram. Soc. 2014, 34, 2933–2939. [Google Scholar] [CrossRef]
  45. Bobić, J.D.; Ivanov, M.; Ilić, N.I.; Dzunuzović, A.S.; Vijatović Petrović, M.M.; Banys, J.; Ribic, A.; Despotovic, Z.; Stojanovic, B.D. PZT-nickel ferrite and PZT-cobalt ferrite comparative study: Structural, dielectric, ferroelectric and magnetic properties of composite ceramics. Ceram. Int. 2018, 44, 6551–6557. [Google Scholar] [CrossRef]
  46. Nayak, P.; Badapanda, T.; Kumar Singh, A.; Panigrahi, S. An approach for correlating the structural and electrical properties of Zr4+-modified SrBi4Ti4O15/SBT ceramic. RSC Adv. 2017, 7, 16319–16331. [Google Scholar] [CrossRef] [Green Version]
  47. Li, M.; Sinclair, D.C.; West, A.R. Extrinsic origins of the apparent relaxor like behavior in CaCu3Ti4O12 ceramics at high temperatures: A cautionary tale. J. Appl. Phys. 2011, 109, 084106. [Google Scholar] [CrossRef]
  48. Irvine, J.T.; Sinclair, D.C.; West, A.R. Electroceramics: Characterization by impedance spectroscopy. Adv. Mat. 1990, 2, 132–138. [Google Scholar] [CrossRef]
  49. Zhang, M.F.; Wang, Y.; Wang, K.F.; Zhu, J.S.; Liu, J.M. Characterization of oxygen vacancies and their migration in Ba-doped Pb(Zr0.52Ti0.48)O3 ferroelectrics. J. Appl. Phys. 2009, 105, 061639. [Google Scholar] [CrossRef]
  50. Kang, B.S.; Choi, S.K.; Park, C.H. Diffuse dielectric anomaly in perovskite-type ferroelectric oxides in the temperature range of 400–700 °C. J. Appl. Phys. 2003, 9, 1904–1911. [Google Scholar] [CrossRef]
  51. Elissalde, C.; Ravez, J. Ferroelectric ceramics: Defects and dielectric relaxations. J. Mater. Chem. 2001, 11, 1957–1967. [Google Scholar] [CrossRef]
  52. Li, Z.; Fan, H. Polaron relaxation associated with the localized oxygen vacancies in Ba0.85Sr0.15TiO3 ceramics at high temperatures. J. Appl. Phys. 2009, 106, 054102. [Google Scholar] [CrossRef]
  53. Lvovich, V.F. Impedance Spectroscopy, Applications to Electrochemical and Dielectric Phenomena; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2012. [Google Scholar]
  54. Wu, J.; Wang, J. Ferroelectric and impedance behavior of La- and Ti-codoped BiFeO3 thin films. J. Am. Ceram. Soc. 2010, 93, 2795–2803. [Google Scholar] [CrossRef]
  55. Chang, B.Y.; Park, S.M. Electrochemical impedance spectroscopy. Ann. Rev. Anal. Chem. 2010, 3, 207. [Google Scholar] [CrossRef] [PubMed]
  56. Akkopru-Akgun, B.; Marincel, D.M.; Tsuji, K.; Bayer, T.J.M.; Randall, C.A.; Lanagan, M.T.; Trolier-McKinstry, S. Thermally stimulated depolarization current measurements on degraded lead zirconate titanate films. J. Am. Ceram. Soc. 2021, 104, 5270–5280. [Google Scholar] [CrossRef]
Figure 1. SEM analysis of P-F multiferroic composites developed using a standard SB mode (top) and BSE technique (bottom): P20-F80 (a,a′), P40-F60 (b,b′), P60-F40 (c,c′), P80-F20 (d,d′), P85-F15 (e,e′), and P90-F10 (f,f′).
Figure 1. SEM analysis of P-F multiferroic composites developed using a standard SB mode (top) and BSE technique (bottom): P20-F80 (a,a′), P40-F60 (b,b′), P60-F40 (c,c′), P80-F20 (d,d′), P85-F15 (e,e′), and P90-F10 (f,f′).
Applsci 13 08854 g001
Figure 2. Grain size distribution of the ferroelectric grains (af) and magnetic grains (a′f′) for the P−F composites: P20-F80 (a,a′), P40-F60 (b,b′), P60-F40 (c,c′), P80-F20 (d,d′), P85-F15 (e,e′), and P90-F10 (f,f′).
Figure 2. Grain size distribution of the ferroelectric grains (af) and magnetic grains (a′f′) for the P−F composites: P20-F80 (a,a′), P40-F60 (b,b′), P60-F40 (c,c′), P80-F20 (d,d′), P85-F15 (e,e′), and P90-F10 (f,f′).
Applsci 13 08854 g002
Figure 3. EDS analysis of the P-F multiferroic composites: P20-F80 (a), P40-F60 (b), P60-F40 (c), P80-F20 (d), P85-F15 (e), P90-F10 (f), and EPMA maps (g) for the P90-F10 sample.
Figure 3. EDS analysis of the P-F multiferroic composites: P20-F80 (a), P40-F60 (b), P60-F40 (c), P80-F20 (d), P85-F15 (e), P90-F10 (f), and EPMA maps (g) for the P90-F10 sample.
Applsci 13 08854 g003
Figure 4. X-ray diffraction patterns for the P-F multiferroic composites (according to [37]).
Figure 4. X-ray diffraction patterns for the P-F multiferroic composites (according to [37]).
Applsci 13 08854 g004
Figure 5. Complex impedance plots in temperatures range from 40 to 400 °C for multiferroic composite samples: P20-F80 (a), P40-F60 (b), P60-F40 (c), P80-F20 (d), P85-F15 (e), and P90-F10 (f).
Figure 5. Complex impedance plots in temperatures range from 40 to 400 °C for multiferroic composite samples: P20-F80 (a), P40-F60 (b), P60-F40 (c), P80-F20 (d), P85-F15 (e), and P90-F10 (f).
Applsci 13 08854 g005
Figure 6. Complex impedance plots in temperature range from 200 to 300 °C for multiferroic composite samples: P20-F80 (a), P40-F60 (b), P60-F40 (c), P80-F20 (d), P85-F15 (e), and P90-F10 (f).
Figure 6. Complex impedance plots in temperature range from 200 to 300 °C for multiferroic composite samples: P20-F80 (a), P40-F60 (b), P60-F40 (c), P80-F20 (d), P85-F15 (e), and P90-F10 (f).
Applsci 13 08854 g006
Figure 7. The normalized imaginary parts Z″/Zmax of impedance as a function of frequency in temperature range from 200 to 300 °C for multiferroic composite samples: P20-F80 (a), P40-F60 (b), P60-F40 (c), P80-F20 (d), P85-F15 (e), and P90-F10 (f).
Figure 7. The normalized imaginary parts Z″/Zmax of impedance as a function of frequency in temperature range from 200 to 300 °C for multiferroic composite samples: P20-F80 (a), P40-F60 (b), P60-F40 (c), P80-F20 (d), P85-F15 (e), and P90-F10 (f).
Applsci 13 08854 g007
Figure 8. Dependence of the activation-energy values (a) and compositional tuning (b) with the ferrite concentration for obtained P-F composite samples.
Figure 8. Dependence of the activation-energy values (a) and compositional tuning (b) with the ferrite concentration for obtained P-F composite samples.
Applsci 13 08854 g008
Table 1. The calculations of theoretical and experimental results of oxides in the composition.
Table 1. The calculations of theoretical and experimental results of oxides in the composition.
P20-F80P40-F60P60-F40P80-F20P85-F15P90-F10
t (%)ex (%)t (%)ex (%)t (%)ex (%)t (%)ex (%)t (%)ex (%)t (%)ex (%)
TiO23.0652.445.7465.348.1098.0510.20810.0710.69710.4711.17310.92
MnO20.0070.090.0130.120.0180.150.0230.170.0240.120.0250.17
Fe2O350.13355.0535.23839.3522.10321.9310.43510.137.7187.635.0764.91
NiO11.725138.24110.875.1695.082.4402.311.8051.851.1871.13
ZnO12.77913.68.98211.055.6345.782.6602.631.9671.781.2941.08
ZrO24.9232.859.2277.0313.02212.9516.39316.2317.17816.8517.94217.58
Nb2O50.0100.260.0200.310.0280.770.0350.970.0361.010.0381.02
PbO16.99312.3431.85025.3944.95144.2756.59056.3959.29958.9761.93561.42
Bi2O30.3660.370.6850.540.9671.0201.2181.101.2761.321.3331.77
Table 2. Inverse of the peak frequency (relaxation times) as a function of P-F composition.
Table 2. Inverse of the peak frequency (relaxation times) as a function of P-F composition.
CompositionER1 for I PeakER2 for II Peak
P20-F800.18 ± 0.1 eV0.17 ± 0.1 eV
P40-F600.16 ± 0.1 eV0.15 ± 0.1 eV
P60-F400.18 ± 0.1 eV0.17 ± 0.1 eV
P80-F200.17 ± 0.1 eV-
P85-F150.15 ± 0.1 eV-
P90-F100.16 ± 0.1 eV-
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kozielski, L.; Bochenek, D.; Clemens, F.; Sebastian, T. Magnetoelectric Composites: Engineering for Tunable Filters and Energy Harvesting Applications. Appl. Sci. 2023, 13, 8854. https://doi.org/10.3390/app13158854

AMA Style

Kozielski L, Bochenek D, Clemens F, Sebastian T. Magnetoelectric Composites: Engineering for Tunable Filters and Energy Harvesting Applications. Applied Sciences. 2023; 13(15):8854. https://doi.org/10.3390/app13158854

Chicago/Turabian Style

Kozielski, Lucjan, Dariusz Bochenek, Frank Clemens, and Tutu Sebastian. 2023. "Magnetoelectric Composites: Engineering for Tunable Filters and Energy Harvesting Applications" Applied Sciences 13, no. 15: 8854. https://doi.org/10.3390/app13158854

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