Next Article in Journal / Special Issue
Experimental Investigation of the Magnetoelectric Effect in NdFeB-Driven A-Line Shape Terfenol-D/PZT-5A Structures
Previous Article in Journal
Influence of Synthetic Bone Substitutes on the Anchorage Behavior of Open-Porous Acetabular Cup
Previous Article in Special Issue
Damage from Coexistence of Ferroelectric and Antiferroelectric Domains and Clustering of O Vacancies in PZT: An Elastic and Raman Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Effects of Magnetostrictive Materials on the Magnetoelectric Response of Particulate CZFO/NKNLS Composites

Department of Chemical Engineering, Dong-A University, Busan 49315, Korea
*
Author to whom correspondence should be addressed.
Materials 2019, 12(7), 1053; https://doi.org/10.3390/ma12071053
Submission received: 12 February 2019 / Revised: 27 March 2019 / Accepted: 28 March 2019 / Published: 30 March 2019
(This article belongs to the Special Issue Functional Electroceramics)

Abstract

:
In this study, magnetostrictive powders of CoFe2O4 (CFO) and Zn-substituted CoFe2O4 (CZFO, Zn = 0.1, 0.2) were synthesized in order to decrease the optimal dc magnetic field (Hopt.), which is required to obtain a reliable magnetoelectric (ME) voltage in a 3-0 type particulate composite system. The CFO powders were prepared as a reference via a typical solid solution process. In particular, two types of heterogeneous CZFO powders were prepared via a stepwise solid solution process. Porous-CFO and dense-CFO powders were synthesized by calcination in a box furnace without and with pelletizing, respectively. Then, heterogeneous structures of pCZFO and dCZFO powders were prepared by Zn-substitution on calcined powders of porous-CFO and dense-CFO, respectively. Compared to the CFO powders, the heterogeneous pCZFO and dCZFO powders exhibited maximal magnetic susceptibilities (χmax) at lower Hdc values below ±50 Oe and ±10 Oe, respectively. The Zn substitution effect on the Hdc shift was more dominant in dCZFO than in pCZFO. This might be because the Zn ion could not diffuse into the dense-CFO powder, resulting in a more heterogeneous structure inducing an effective exchange-spring effect. As a result, ME composites consisting of 0.948Na0.5K0.5NbO3–0.052LiSbO3 (NKNLS) with CFO, pCZFO, and dCZFO were found to exhibit Hopt. = 966 Oe (NKNLS-CFO), Hopt. = 689–828 Oe (NKNLS-pCZFO), and Hopt. = 458–481 Oe (NKNLS-dCZFO), respectively. The low values of Hopt. below 500 Oe indicate that the structure of magnetostrictive materials should be considered in order to obtain a minimal Hopt. for high feasibility of ME composites.

Graphical Abstract

1. Introduction

Since the year 2000, magnetoelectric (ME) response has been a topic of interest in the development of energy-harvesters, sensitive magnetic sensors, and magnetically driven memories, or magnetoelectric transducers [1,2,3]. The ME effect is a result of induced piezoelectric effect (electrical effect/mechanical) in a piezoelectric phase by strain transfer of the magnetostrictive effect (mechanical/magnetic) in a magnetostrictive phase [4,5,6,7,8].
ME effect = electric mechanical × mechanical magnetic
However, reliable ME voltage from 3-0 type particulate composites can only be obtained under an optimal dc magnetic field (Hopt.) on the order of over several thousand Oersteds (Oe), which is a serious drawback limiting practical ME applications [9,10]. According to previous studies on particulate ME composites, a maximum ME voltage (αME) was obtained at high values of Hopt. above 1000 Oe from various compositions of Pb(Zr0.52Ti0.48)O3-Ni0.8Zn0.2Fe2O4 (αME = 54.4 mV/cm·Oe at Hopt. = 1000 Oe), BaTiO3-Co0.6Zn0.4Fe1.7Mn0.3O4 (αME = 73 mV/cm·Oe at Hopt. > 2000 Oe), BaTiO3-CoFe2O4 (αME = 17.04 mV/cm·Oe at Hopt. > 15000 Oe), Ba0.85Ca0.15Ti0.9Zr0.1O3-CoFe2O4 (αME = 1.028 mV/cm·Oe at Hopt. > 8000 Oe), and Na0.5Bi0.5TiO3-CoFe2O4 (αME = 0.42 mV/cm·Oe at Hopt. >2500 Oe) [11,12,13,14,15]. Even though lower Hopt. values of 500–1000 Oe were reported when investigating the size effect of magnetostrictive particles in BaTiO3-NiFe1.98O4 (αME = ~252 mV/cm·Oe at Hopt. = 500–1000 Oe), the sintering temperature effect in Pb(Zr0.52Ti0.48)O3-NiCo0.02Cu0.02Mn0.1Fe1.8O4 (αME = 63 mV/cm·Oe at Hopt. = 600 Oe), and the piezoelectric phase effect in Pb(Zr0.52Ti0.48)O3-Ni1−xZnxFe2O4 (αME = 190 mV/cm·Oe at Hopt. = 800 Oe), there is still a need to decrease Hopt. below 100 Oe for a high feasibility of particulate ME composite [16,17,18].
In this study, the structural effects of magnetostrictive materials on ME response was investigated in order to decrease Hopt. values in a particulate ME composite system. In particular, magnetostrictive powders of CoFe2O4 (CFO), Zn-substituted porous-CFO (pCZFO) and Zn-substituted dense-CFO (dCZFO) were respectively prepared to explore structure-dependent hysteretic magnetizations. Then the Hopt. shift in ME response was analyzed in particulate ME composites consisting of each magnetostrictive powder (CFO, pCZFO, and dCZFO) in a 0.948Na0.5K0.5NbO3–0.052LiSbO3 (NKNLS) piezoelectric matrix.

2. Experimental

Figure 1a–c shows a schematic diagram of the experimental procedure based on a solid-solution method to synthesize magnetostrictive powders of CFO, pCZFO, and dCZFO, respectively. As shown in Figure 1a, for preparation of CFO powders, Co3O4 (Sigma-Aldrich, Seoul, Korea, ≥99.5%) and Fe2O3 (Sigma-Aldrich, Seoul, Korea, ≥99.0%) powders were mixed by ball milling for 24 h. The well-mixed and fully dried powders were calcined at 1000 °C for 2 h. The calcined powders were ball-milled for 24 h and then sintered at 1200 °C for 2 h. After crushing and sieving of the sintered powders, CFO powders were selected with a particle size of 24–64 μm. As shown in Figure 1b,c, for preparation of pCZFO and dCZFO powders, Co3O4 (Sigma-Aldrich, Seoul, Korea, ≥99.5%) and Fe2O3 (Sigma-Aldrich, Seoul, Korea, ≥99.0%) powders were mixed by ball milling for 24 h. Then, the well-mixed and fully dried powders were calcined at 1000 °C for 2 h without and with pelletizing at 30 bar pressure, respectively. The calcined CFO powders exhibiting a porous structure (pCFO) and a dense structure (dCFO) were mixed with 0.1 and 0.2 molar ratio of ZnO powders (Sigma-Aldrich, Seoul, Korea, ≥99.0%), respectively. Then the mixed powders were sintered at 1200 °C for 2 h. After crushing and sieving of the sintered powders, pCZFO and dCZFO powders were selected with particle sizes of 24–64 μm.
ME composites were prepared with a 3-0 type particulate structure consisting of the magnetostrictive powders (CFO, pCZFO, and dCZFO, respectively) in a lead-free piezoelectric matrix of NKNLS. For preparation of NKNLS powders, K2CO3 (Sigma-Aldrich, Seoul, Korea, 99%), Na2CO3 (Sigma-Aldrich, Seoul, Korea, 99.5%), Li2CO3 (Sigma-Aldrich, Seoul, Korea, 99%), Nb2O5 (Sigma-Aldrich, Seoul, Korea, 99.9%), and Sb2O5 (Sigma-Aldrich, Seoul, Korea, 99%) powders were mixed by ball milling for 24 h. Then, the well-mixed and fully dried powders were calcined at 880 °C for 2 h. After sintering at 1050 °C for 2 h of CFO-NKNLS, pCZFO-NKNLS, and dCZFO-NKNLS pellets with a magnetostrictive/piezoelectric weight ratio of 0.1, disk-type ME composites were prepared with a thickness of 1 mm and a diameter of 13 mm. The ME composites were poled in silicone oil at room temperature by applying a dc field of 3 kV/mm for 30 min.
Crystal structures were investigated by X-ray diffraction (XRD; Miniflex600, RIGAKU, Tokyo, Japan) with CuKα (λ = 1.5406 Å) radiation. The surface morphology was investigated by scanning electron microscopy (SEM; JEOL-6700F, Tokyo, Japan). Hysteretic magnetization curves were characterized by vibrating sample magnetometry (VSM; Model 7404, Lakeshore, CA, USA). Piezoelectric constants were measured by an APC YE 2730A d33 meter (APC Inc., Mackeyville, PA, USA). ME voltages were measured by applying an Hac of 1 Oe at an off-resonance frequency, f, of 1 kHz using a lock-in amplifier (SR860, Stanford Research Systems Inc., Sunnyvale, CA, USA) [19,20]. As shown in Figure 1d, using the lock-in amplifier a calculated ac current was applied to a Helmholtz coil to induce an Hac of 1 Oe with an off-resonance frequency of 1 kHz. Then, an Hdc of ±1000 Oe was applied to the ME samples using an electromagnet to obtain reliable ME voltages. Output ac voltage (Vac) from the ME samples was measured by the lock-in amplifier.

3. Results and Discussion

Crystal structures of the magnetostrictive CFO, pCZFO (Zn = 0.1, Zn = 0.2), and dCZFO (Zn = 0.1, Zn = 0.2) powders were investigated from XRD patterns. As shown in Figure 2a, all magnetostrictive powders were found to exhibit XRD peaks of (220), (311), (222), (400), (422), (511), and (440) representing a spinel structure of AB2O4 (JCPDS card No. 22-1086) [21,22]. Even though no noticeable peak shift in the XRD patterns was observed over a wide 2θ range after Zn substitution of 0.1 and 0.2 molar ratio on the porous-CFO and dense-CFO powders, a major shift of the (311) peak at 2θ = 35.5° towards a lower angle by Zn substitution was observed in the XRD patterns at a narrow 2θ range, as shown in Figure 2b. Bragg’s Law can be used to calculate a lattice constant using the equation:
a2 = λ2(h2 + k2 + l2)1/2/4sin2θ
where a is the lattice constant, λ is the wavelength of CuKα radiation, and h, k, and l are the Miller indices. As the (311) peak shifts to a lower angle by Zn substitution, the lattice constant increases due to a decrease in the value of sin θ. With respect to the ionic radius, the pCZFO and dCZFO powders were found to exhibit an increased lattice constant compared to CFO powders because Zn2+ (0.82 Å) has a larger ionic radius than Co2+ (0.78 Å), which is replaced by Zn2+ [23,24,25].
In terms of Zn substitution in the porous-CFO and dense-CFO powders, magnetic properties of saturation magnetization (Ms), remanent magnetization (Mr), coercive field (Hc), and magnetic susceptibility (χ = dM/dH) were investigated, as shown in Figure 3 and Table 1. Compared to the CFO powders, the pCZFO and dCZFO powders were found to exhibit enhanced Ms with decreased Hc, as shown in Table 1. The enhanced values of Ms demonstrate that the addition of Zn2+ ions causes a migration of Fe3+ ions from a tetrahedral site to an octahedral site, which causes an increase of the total magnetic moment by reducing the net magnetic moment in the tetrahedral site. Furthermore, decreased values of Hc illustrate that grain growth by Zn substitution causes an increase of the domain wall number, resulting in large grain size, which requires less energy for spin rotation [26,27]. As shown in Figure 3b,e, stepped demagnetization behavior is shown by pCZFO with Zn = 0.2 and dCZFO with Zn = 0.1 and 0.2, which might be caused by the exchange-spring effect derived from the interplay of two uniquely characteristic phases [28,29,30]. From the result, it is noted that dCZFO possesses a sufficient exchange-spring effect based on high interaction between two magnetostrictive phases even though the Zn substitution of 0.1 is low in the dense-CFO powders. As shown in Figure 3c,f, the pCZFO and dCZFO powders were found to exhibit higher χmax of 0.22–0.42 emu/g·Oe at lower values of Hdc below ±50 Oe, compared to χmax of 0.05 emu/g·Oe at an Hdc below ±200 Oe from the CFO powders. In particular, the χmax values of dCZFO were obtained at very low values of Hdc below ±10 Oe, which are induced by prominent stepped demagnetization behavior.
To investigate structure-dependent ME responses, particulate ME composites were prepared with compositions of CFO-NKNLS, pCZFO-NKNLS (Zn = 0.1, 0.2), and dCZFO-NKNLS (Zn = 0.1, 0.2). From the XRD patterns, as shown in Figure 4, perovskite (ABO3) and spinel (AB2O4) crystal structures were confirmed as piezoelectric and magnetostrictive phases, respectively. Even though sintering was conducted at 1050 °C for 2 h, all ME composites were found to exhibit stable crystal structures without any trace of secondary phase. In particular, a peak split at 2θ = 45–46° representing a tetragonal phase was maintained during the high temperature sintering. Therefore, the ME composites were found to exhibit a piezoelectric charge constant (d33) of 55–60 pC/N after sample poling.
From the particulate composites of CFO-NKNLS, pCZFO-NKNLS (Zn = 0.1, 0.2), and dCZFO-NKNLS (Zn = 0.1, 0.2), ME voltage (αME) and Hopt. were investigated while applying Hac = 1 Oe at f = 1 kHz by sweeping Hdc of ±1000 Oe, as shown in Figure 5 and Table 2. The CFO-NKNLS composites were found to exhibit a maximum αME = 140 μV/cm·Oe at Hopt. = 966 Oe. Even though a decreased Hopt. value of 689–828 Oe was obtained from pCZFO-NKNLS as shown in Figure 5a, there was not a sufficient Hopt. shift due to its weak behavior of stepped demagnetization. On the other hand, the dCZFO-NKNLS composites were found to exhibit remarkable Hopt. values of 458–481 Oe as shown in Figure 5b, which are lower Hopt. values than any reported particulate ME composites so far. As a result, the structural effect of magnetostrictive powders on Hopt. shift is clearly shown between the heterogeneous pCZFO and dCZFO powders. Although the obtained Hopt. value of 458 Oe from dCZFO-NKNLS is higher than 100 Oe, this study can serve to minimize a required Hopt. by complexation with previous studies for high feasibility of particulate ME composites.

4. Conclusions

In this study, magnetostrictive powders of CFO, pCZFO (Zn = 0.1, 0.2) and dCZFO (Zn = 0.1, 0.2) were prepared to produce low values of Hopt., which is required to obtain a reliable ME voltage in a 3-0 type particulate composite system. Compared to the CFO powders (χmax = 0.05 emu/g·Oe at Hdc below ±200 Oe), the pCZFO and dCZFO powders were found to exhibit higher χmax of 0.22–0.42 emu/g·Oe at lower Hdc values below ±50 Oe and ±10 Oe, respectively. The NKNLS-based ME composites consisting of CFO, pCZFO, dCZFO, respectively were found to exhibit Hopt. = 966 Oe (NKNLS-CFO), Hopt. = 689–828 Oe (NKNLS-pCZFO), and Hopt. = 458–481 Oe (NKNLS-dCZFO). The results illustrate that a low Hopt. value of 458 Oe was obtained from the effective stepped demagnetization behavior of dCZFO (Zn = 0.2), which was induced by a structural effect in a heterogeneous magnetostrictive phase.

Author Contributions

Conceptualization, M.H.C. and S.C.Y.; methodology, M.H.C. and S.C.Y.; validation, M.H.C., K.K. and S.C.Y.; formal analysis, M.H.C.; investigation, M.H.C. and K.K.; resources, S.C.Y.; data curation, M.H.C. and K.K.; writing—original draft preparation, M.H.C.; writing—review and editing, S.C.Y.; visualization, M.H.C. and S.C.Y.; supervision, S.C.Y.; project administration, S.C.Y.; funding acquisition, S.C.Y.

Funding

This research was financially supported by the Dong-A University research fund.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cheng, Y.; Peng, B.; Hu, Z.; Zhou, Z.; Liu, M. Recent development and status of magnetoelectric materials and devices. Phys. Lett. A 2018, 382, 3018–3025. [Google Scholar] [CrossRef]
  2. Palneedi, H.; Annapureddy, V.; Priya, S.; Ryu, J. Status and Perspectives of Multiferroic Magnetoelectric Composite Materials and Applications. Actuators 2016, 5, 9. [Google Scholar] [CrossRef]
  3. Ren, Y.; Ouyang, J.; Wang, W.; Wu, X.; Yang, X.; Zhang, Y.; Chen, S. Rotating Magnetoelectric Sensor for DC Magnetic Field Measurement. IEEE Trans. Magn. 2018, 54, 6001203. [Google Scholar]
  4. Yang, S.-C.; Ahn, C.-W.; Cho, K.-H.; Priya, S. Self-Bias Response of Lead-Free (1−x)[0.948 K0.5Na0.5NbO3–0.052 LiSbO3]-xNi0.8Zn0.2Fe2O4-Nickel Magnetoelectric Laminate Composites. J. Am. Ceram. Soc. 2011, 94, 3889–3899. [Google Scholar] [CrossRef]
  5. Eerenstein, W.; Mathur, N.D.; Scott, J.F. Multiferroic and magnetoelectric materials. Nature 2006, 442, 759–765. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, Y.; Li, J.; Viehland, D. Magnetoelectrics for magnetic sensor applications: Status, challenges and perspectives. Mater. Today 2014, 17, 269–275. [Google Scholar] [CrossRef]
  7. Fiebig, M. Revival of the magnetoelectric effect. J. Phys. D-Appl. Phys. 2005, 38, R123–R152. [Google Scholar] [CrossRef]
  8. Nan, C.W.; Bichurin, M.I.; Dong, S.X.; Viehland, D.; Srinivasan, G. Multiferroic magnetoelectric composites: Historical perspective, status, and future directions. J. Appl. Phys. 2008, 103, 1. [Google Scholar] [CrossRef]
  9. Li, T.; Ma, D.; Li, K.; Hu, Z. Self-biased magnetoelectric coupling effect in the layered La0.7Sr0.3MnO3/BaTiO3/La0.7Sr0.3MnO3 multiferroic heterostructure. J. Alloy. Compd. 2018, 747, 558–562. [Google Scholar] [CrossRef]
  10. An, F.; Zhong, G.; Zhu, Q.; Huang, Y.; Yang, Y.; Xie, S. Synthesis and mechanical properties characterization of multiferroic BiFeO3-CoFe2O4 composite nanofibers. Ceram. Int. 2018, 44, 11617–11621. [Google Scholar] [CrossRef]
  11. Islam, R.A.; Priya, S. Effect of piezoelectric grain size on magnetoelectric coefficient of Pb(Zr0.52Ti0.48)O3-Ni0.8Zn0.2Fe2O4 particulate composites. J. Mater. Sci. 2008, 43, 3560–3568. [Google Scholar] [CrossRef]
  12. Gupta, A.; Chatterjee, R. Dielectric and magnetoelectric properties of BaTiO3-Co0.6Zn0.4Fe1.7Mn0.3O4 composite. J. Eur. Ceram. Soc. 2013, 33, 1017–1022. [Google Scholar] [CrossRef]
  13. Nie, J.W.; Xu, G.Y.; Yang, Y.; Cheng, C.W. Strong magnetoelectric coupling in CoFe2O4-BaTiO3 composites prepared by molten-salt synthesis method. Mater. Chem. Phys. 2009, 115, 400–403. [Google Scholar] [CrossRef]
  14. Negi, N.S.; Kumar, R.; Sharma, H.; Shah, J.; Kotnala, R.K. Structural, multiferroic, dielectric and magnetoelectric properties of (1-x) Ba0.85Ca0.15Ti0.90Zr0.10O3-(x)CoFe2O4 lead-free composites. J. Magn. Magn. Mater. 2018, 456, 292–299. [Google Scholar] [CrossRef]
  15. Walther, T.; Straube, U.; Koferstein, R.; Ebbinghaus, S.G. Hysteretic magnetoelectric behavior of CoFe2O4-BaTiO3 composites prepared by reductive sintering and reoxidation. J. Mater. Chem. C 2016, 4, 4792–4799. [Google Scholar] [CrossRef]
  16. Sreenivasulu, G.; Babu, V.H.; Markandeyulu, G.; Murty, B.S. Magnetoelectric effect of (100-x)BaTiO3-(x)NiFe1.98O4 (x = 20–80 wt %) particulate nanocomposites. Appl. Phys. Lett. 2009, 94, 112902. [Google Scholar] [CrossRef]
  17. Ryu, J.; Carazo, A.V.; Uchino, K.; Kim, H.E. Piezoelectric and magnetoelectric properties of Lead Zirconate Titanate/Ni-Ferrite particulate composites. J. Electroceram. 2001, 7, 17–24. [Google Scholar] [CrossRef]
  18. Islam, R.A.; Viehland, D.; Priya, S. Doping effect on magnetoelectric coefficient of Pb(Zr052Ti0.48)O3–Ni(1−x)ZnxFe2O4 particulate. J. Mater. Sci. 2008, 43, 1497–1500. [Google Scholar] [CrossRef]
  19. Yang, S.-C.; Kumar, A.; Petkov, V.; Priya, S. Room-temperature magnetoelectric coupling in single-phase BaTiO3-BiFeO3 system. J. Appl. Phys. 2013, 113, 144101. [Google Scholar] [CrossRef]
  20. Shovon, O.G.; Rahaman, M.D.; Tahsin, S.; Hossain, A.K.M.A. Synthesis and characterization of (100-x) Ba0.82Sr0.03Ca0.15Zr0.10Ti0.90O3 + (x) Mg0.25Cu0.25Zn0.5Mn0.05Fe1.95O4 composites with improved magnetoelectric voltage coefficient. J. Alloy. Compd. 2018, 735, 291–311. [Google Scholar] [CrossRef]
  21. Allaedini, G.; Tasirin, S.M.; Aminayi, P. Magnetic properties of cobalt ferrite synthesized by hydrothermal method. Int. Nano Lett. 2015, 5, 183–186. [Google Scholar] [CrossRef] [Green Version]
  22. Ben Ali, M.; El Maalam, K.; El Moussaoui, H.; Mounkachi, O.; Hamedoun, M.; Masrour, R.; Hlil, E.K.; Benyoussef, A. Effect of zinc concentration on the structural and magnetic properties of mixed Co–Zn ferrites nanoparticles synthesized by sol/gel method. J. Magn. Magn. Mater. 2016, 398, 20–25. [Google Scholar] [CrossRef]
  23. Ansari, S.M.; Sinha, B.B.; Pai, K.R.; Bhat, S.K.; Ma, Y.-R.; Sen, D.; Kolekar, Y.D.; Ramana, C.V. Controlled surface/interface structure and spin enabled superior properties and biocompatibility of cobalt ferrite nanoparticles. Appl. Surf. Sci. 2018, 459, 788–801. [Google Scholar] [CrossRef]
  24. Köseoğlu, Y.; Baykal, A.; Gözüak, F.; Kavas, H. Structural and magnetic properties of CoxZn1−xFe2O4 nanocrystals synthesized by microwave method. Polyhedron 2009, 28, 2887–2892. [Google Scholar] [CrossRef]
  25. Vaidyanathan, G.; Sendhilnathan, S. Characterization of Co1−xZnxFe2O4 nanoparticles synthesized by co-precipitation method. Phys. B 2008, 403, 2157–2167. [Google Scholar] [CrossRef]
  26. Anjum, S.; Khurram, R.; Bashir, F.; Nazli, H. Fabrication and Investigation of Structural, Magnetic and Dielectrical Properties of Zn Substituted Co-ferrites. Mater. Today Proc. 2015, 2, 5515–5521. [Google Scholar] [CrossRef]
  27. Praveena, K.; Sadhana, K.; Liu, H.-L.; Murthy, S.R. Effect of Zn substitution on structural, dielectric and magnetic properties of nanocrystalline Co1−xZnxFe2O4 for potential high density recording media. J. Mater. Sci.-Mater. Electron. 2016, 27, 12680–12690. [Google Scholar] [CrossRef]
  28. Bill, A.; Braun, H.B. Magnetic properties of exchange springs. J. Magn. Magn. Mater. 2004, 272–276, 1266–1267. [Google Scholar] [CrossRef]
  29. Chithra, M.; Anumol, C.N.; Sahu, B.; Sahoo, S.C. Exchange spring like magnetic behavior in cobalt ferrite nanoparticles. J. Magn. Magn. Mater. 2016, 401, 1–8. [Google Scholar] [CrossRef]
  30. Lavorato, G.; Winkler, E.; Rivas-Murias, B.; Rivadulla, F. Thickness dependence of exchange coupling in epitaxial Fe3O4/CoFe2O4soft/ard magnetic bilayers. Phys. Rev. B 2016, 94, 054405. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of experimental procedure based on solid-solution synthesis for (a) CoFe2O4 (CFO), (b) Zn-substituted porous-CoFe2O4 (pCZFO), and (c) Zn-substituted dense-CoFe2O4 (dCZFO) powders. (d) Schematic diagram of magnetoelectric measurement set up.
Figure 1. Schematic diagram of experimental procedure based on solid-solution synthesis for (a) CoFe2O4 (CFO), (b) Zn-substituted porous-CoFe2O4 (pCZFO), and (c) Zn-substituted dense-CoFe2O4 (dCZFO) powders. (d) Schematic diagram of magnetoelectric measurement set up.
Materials 12 01053 g001
Figure 2. XRD patterns of CFO, pCZFO (Zn = 0.1, 0.2) and dCZFO (Zn = 0.1, 0.2) powders; (a) wide range 2θ of 20–70° and (b) narrow range 2θ of 35.0–35.8°.
Figure 2. XRD patterns of CFO, pCZFO (Zn = 0.1, 0.2) and dCZFO (Zn = 0.1, 0.2) powders; (a) wide range 2θ of 20–70° and (b) narrow range 2θ of 35.0–35.8°.
Materials 12 01053 g002
Figure 3. Hysteretic magnetization curves at (a,d) a wide Hdc range of ±10 kOe and (b,e) a narrow Hdc range of ±1 kOe, (c,f) magnetic susceptibilities (χ) of CFO, pCZFO (Zn = 0.1, 0.2) and dCZFO (Zn = 0.1, 0.2) powders.
Figure 3. Hysteretic magnetization curves at (a,d) a wide Hdc range of ±10 kOe and (b,e) a narrow Hdc range of ±1 kOe, (c,f) magnetic susceptibilities (χ) of CFO, pCZFO (Zn = 0.1, 0.2) and dCZFO (Zn = 0.1, 0.2) powders.
Materials 12 01053 g003
Figure 4. XRD patterns of magnetoelectric (ME) particulate composites consisting of a piezoelectric phase of NKNLS and magnetostrictive phases of CFO, pCZFO (Zn = 0.1, 0.2) and dCZFO (Zn = 0.1, 0.2).
Figure 4. XRD patterns of magnetoelectric (ME) particulate composites consisting of a piezoelectric phase of NKNLS and magnetostrictive phases of CFO, pCZFO (Zn = 0.1, 0.2) and dCZFO (Zn = 0.1, 0.2).
Materials 12 01053 g004
Figure 5. ME voltage of particulate composites consisting of a piezoelectric phase of NKNLS and magnetostrictive phases of (a) CFO and pCZFO (Zn = 0.1, 0.2) and (b) CFO and dCZFO (Zn = 0.1, 0.2).
Figure 5. ME voltage of particulate composites consisting of a piezoelectric phase of NKNLS and magnetostrictive phases of (a) CFO and pCZFO (Zn = 0.1, 0.2) and (b) CFO and dCZFO (Zn = 0.1, 0.2).
Materials 12 01053 g005
Table 1. Magnetic properties of CoFe2O4 (CFO), Zn-substituted porous-CoFe2O4 (pCZFO) and Zn-substituted dense-CoFe2O4 (dCZFO) powders; saturation magnetization (Ms), remanent magnetization (Mr), coercive field (Hc), and magnetic susceptibility (χ = dM/dH).
Table 1. Magnetic properties of CoFe2O4 (CFO), Zn-substituted porous-CoFe2O4 (pCZFO) and Zn-substituted dense-CoFe2O4 (dCZFO) powders; saturation magnetization (Ms), remanent magnetization (Mr), coercive field (Hc), and magnetic susceptibility (χ = dM/dH).
Magnetostrictive PowdersZn RatioSaturation MagnetizationRemanant MagnetizationCoercive FieldMagnetic Susceptibility
Ms (emu/g)Mr (emu/g)Hc (Oe)χmax (emu/g·Oe)
CFOZn = 074.5 ± 0.7516.8 ± 0.17366.2 ± 3.660.05
pCZFOZn = 0.177.1 ± 0.7718.5 ± 0.19101.6 ± 1.020.22
Zn = 0.275.9 ± 0.760.5 ± 0.012.4 ± 0.020.42
dCZFOZn = 0.186.3 ± 0.869.3 ± 0.0936.2 ± 0.360.34
Zn = 0.282.6 ± 0.832.3 ± 0.0210.8 ± 0.110.35
Table 2. Magnetoelectric (ME) responses of CFO-NKNLS, pCZFO-NKNLS, and dCZFO-NKNLS composites; optimal magnetic field (Hopt.) and ME voltage (αME).
Table 2. Magnetoelectric (ME) responses of CFO-NKNLS, pCZFO-NKNLS, and dCZFO-NKNLS composites; optimal magnetic field (Hopt.) and ME voltage (αME).
Magnetoelectric CompositesZn RatioOptimal Magnetic FieldMagnetoelectric Voltage
Hopt. (Oe)αME (μV/cm·Oe)
CFO-NKNLSZn = 0966140 ± 21.0
pCZFO-NKNLSZn = 0.1689130 ± 19.5
Zn = 0.2828179 ± 26.9
dCZFO-NKNLSZn = 0.1481228 ± 34.2
Zn = 0.2458184 ± 27.6

Share and Cite

MDPI and ACS Style

Choi, M.H.; Ko, K.; Yang, S.C. Structural Effects of Magnetostrictive Materials on the Magnetoelectric Response of Particulate CZFO/NKNLS Composites. Materials 2019, 12, 1053. https://doi.org/10.3390/ma12071053

AMA Style

Choi MH, Ko K, Yang SC. Structural Effects of Magnetostrictive Materials on the Magnetoelectric Response of Particulate CZFO/NKNLS Composites. Materials. 2019; 12(7):1053. https://doi.org/10.3390/ma12071053

Chicago/Turabian Style

Choi, Moon Hyeok, Kyujin Ko, and Su Chul Yang. 2019. "Structural Effects of Magnetostrictive Materials on the Magnetoelectric Response of Particulate CZFO/NKNLS Composites" Materials 12, no. 7: 1053. https://doi.org/10.3390/ma12071053

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