Multiferroic materials, in which ferroelectric orders coexist with ferromagnetic (or antiferromagnetic) orders, have been brought into focus due to their potential applications in sensors, transducers, and memories.[
Chinese Physics B, Volume. 29, Issue 8, (2020)
Enhanced ferromagnetism and magnetoelectric response in quenched BiFeO3-based ceramics
The piezoelectric, ferromagnetism, and magnetoelectric response of BiFeO3–BaTiO3 ceramics with the compositions around the morphotropic phase boundary (MPB) of the solid solution are systematically investigated after the ceramics have been quenched from a high temperature. We find that the ferromagnetism of the quenched ceramics is greatly enhanced. An enhanced piezoelectric response d33 larger than 200 pC/N, which could be sustained up to 350 °C, is measured. As a result of enhanced ferromagnetism and piezoelectric response, a large magnetoelectric response ~ 1.3 V/cm·Oe (1 Oe = 79.5775 A·m-1) is obtained near the mechanical resonance frequency of the quenched ceramic samples. Our research also shows that in addition to the ferromagnetism and piezoelectric response, the mechanical quality factor is another important parameter to achieve high magnetoelectric response because the physical effects are coupled through mechanical interaction in BiFeO3-based materials. Our work suggests that quenching is an effective approach to enhancing the magnetoelectric response of BiFeO3-based materials and the materials belong to single-phase multiferroic materials with high magnetoelectric response.
1. Introduction
Multiferroic materials, in which ferroelectric orders coexist with ferromagnetic (or antiferromagnetic) orders, have been brought into focus due to their potential applications in sensors, transducers, and memories.[
The magnetoelectric (ME) response is the generation of electric polarization P (or magnetization M) upon applying a magnetic field H (or electric field E).[
Although the effect of quenching on the ferroelectric and piezoelectric properties of BFO-based materials are frequently studied, its effect on the ferromagnetism and ME response has seldom been reported. In this work, the piezoelectric response, ferromagnetism, and ME response of the quenched BFO–BTO ceramics with compositions near MPB are systematically investigated. We show that the quenching process not only improves the piezoelectric response, but also enhances ferromagnetism and ME response. A large ME response > 1 V/cm·Oe is achieved in the BFO–BTO ceramics at mechanical resonance frequency. We also find that the highest ME response is not measured in compositions with the highest piezoelectric response nor with the strongest ferromagnetism. We propose that the mechanical quality factor should play an important role in generating the high ME response in BFO–BTO ceramics.
2. Experimental procedure
The (1−x)BiFeO3–xBaTiO3 ((1−x)BFO–xBTO) ceramic samples with the compositions in a range from x = 0.25 to x = 0.4 were prepared by using the conventional solid-state reaction method. The Bi2O3 (purity 99.9%), BaCO3 (purity 99.9%), Fe2O3 (purity 99.9%), TiO2 (purity 98.0%) were used as starting materials. All the raw materials are weighed according to stoichiometric ratio of the ceramics. Alcohol was added into the raw materials followed by ball milling. The mixture was calcined at 850 °C for 1 h and then at 940 °C for 1 h in sequence. The powder was pressed into disk samples by using polyvinyl alcohol as a binder. The wafers were heated at 600 °C for 3 h to remove the binder and then sintered at 1000 °C–1030 °C for 20 h. The sintered samples were abraded to a thickness of 0.4 mm by using sand paper. The abraded samples were treated at 850 °C for 30 min and then quickly quenched in the air. After being quenched, the ceramic wafers were cut into ceramic bars with the size of 20 mm × 3 mm × 0.4 mm by using a wire cutting machine (STX-202A, Kejing auto-instrument Co., Ltd., Shenyang). The crystal structures of (1−x)BFO–xBTO ceramics were measured by x-ray diffraction (XRD) through using Rigaku Smartlab diffractometer (Rigaku, Tokyo, Japan) The microstructure for each of the ceramic samples was examined by using a scanning electron microscope (SEM, Sirion200, FEI, USA).
Gold electrodes were deposited on the samples for electrical tests by using a sputter coater (EMS150T, Electron Microscopy Sciences, Hatfield, PA, USA). A quasi-static d33 meter (ZJ-6A, Institute of Acoustics, CAS, Beijing, China) was used to measure the piezoelectric coefficient d33 for each of the poled (1−x)BFO–xBTO samples. The poling process was performed at an electric field of 5 kV/mm at 120 °C for 15 min. The dielectric properties of (1−x)BFO–xBTO ceramics were measured by an LCR meter (model E4980, Agilent Technology, Santa Clara, CA, USA). The polarization–electric field (P–E) hysteresis loops were measured by using a modified Sawyer–Tower circuit (Polyktech, State College, USA). The vibrating sample magnetometer (SQUID-VSM, Quantum design, USA) was used to measure the magnetic hysteresis loops of (1−x)BFO–xBTO samples. The ME response was determined by a commercial ME measurement system (Super ME, Quantum design, USA). The magnetoelectric response of the BFO–BTO ceramics was measured by using the same procedure as that in Refs. [34,35]. During the measurement, a direct current (DC) magnetic field was applied to the ceramic samples. At the same time, a small AC magnetic field (< 2 Oe) was applied to the samples to attain the magnetoelectric coefficient at this specific DC field. The DC magnetic field can be varied to obtain the coefficients under different DC fields. The impedance spectra of the poled (1−x)BFO–xBTO ceramic samples were measured by an impedance analyzer (4294A, Agilent, Santa Clara, CA) and the mechanical quality factor was obtained from the impedance spectra.[
3. Results and discussion
Figure 1(a) shows the XRD patterns of (1−x)BFO–xBTO ceramic samples with x in a range from 0.25 to 0.4. All the compositions have a perovskite structure without impurity phase according to the XRD patterns.[
Figure 1.(a) XRD patterns of (1−
Figure 2(a) shows the temperature dependence of the weak-field dielectric constant and loss of (1−x)BFO–xBTO ceramic samples. The dielectric maximum temperature (Tm) can be regarded as the Curie temperature of the material. Consistent with prior studies, Tm gradually moves to a lower temperature with the increase of BTO content, indicating the lowering of Curie temperature of the solid solution with more BTO content in the composition.[
Figure 2.(a) Temperature-dependent dielectric constants and dielectric losses of (1−
The piezoelectric response is an important parameter for generating the strong ME response. The variations of piezoelectric coefficient d33 with composition of the quenched BFO–BTO ceramics are shown in Fig. 3(a). Like the observation in PZT, the maximum d33 is measured to be 215 pC/N in the quenched 0.7BFO–0.3BTO ceramics, a composition near the MPB of the solid solution. In Fig. 3(a), the piezoelectric response of the BFO–BTO ceramics without being subjected to the quenching process is also shown. Consistent with prior studies, the quenching process significantly enhances the piezoelectric response. From Fig. 2(a), we can see that the Tm of 0.7BFO–0.3BTO ceramics with the composition having the highest d33 among the BFO–BTO ceramics that we studied (Fig. 3(a)), is ∼ 500 °C, which is much higher than the Tc of PZT ceramics. It implies that the piezoelectric response of the ceramics can be sustained at a temperature much higher than that of the PZT ceramics. The temperature-dependent d33 of 0.7BFO–0.3BTO ceramics is measured and the result is shown in Fig. 3(b). The d33 of the ceramics is slightly dependent on the temperature below 350 °C, suggesting that BFO–BTO ceramics is a class of lead-free piezoelectric material with a high piezoelectric response, which promises to possess high-temperature applications.
Figure 3.(a) Piezoelectric coefficient
Figure 4(a) shows the magnetic hysteresis (M–H) curves for the quenched (1−x)BFO–xBTO ceramics with the compositions near the MPB. The changes of magnetization with magnetic field for four different compositions are shown in Fig. 4(a), where the inset displays the residual magnetization and coercove field varying with composition. The remnant magnetization increases first, reaching the maximum value when x = 0.3, and decreases then. The variation of magnetic properties with compositions can be attributed to the change of spiral modulated spin structure existing in BFO after incorporating BTO into BFO. The mechanism proposed for the observation of ferromagnetism in BFO–BTO solid solutions was reported in Ref. [16]. We notice that the magnetic properties in the quenched BFO–BTO ceramics are different from those reported in Ref. [16], which can be attributed to the difference in preparation conditions between the studies. Although the magnetic properties of BFO–BTO solid solutions have been widely studied, their magnetic responses in different reports[
Figure 4.(a) Magnetic hysteresis curves of quenched (1−
Quenching is found to have a significant effect on the ferromagnetism of the BFO–BTO ceramics. Figure 4(b) shows M–H curves for 0.7BFO–0.3BTO ceramics before and after quenching. An enhanced ferromagnetism is observed in air-quenched sample compared with the sample without being subjected to the quenching process. The remnant magnetization increases from 0.1263 emu/g to 0.2248 emu/g after quenching. The enhancement of ferromagnetism of the quenched BFO–BTO is not clear and needs further studying. Because the stress enhanced ferromagnetism has been observed in BFO-based thin film,[
The enhanced piezoelectric and magnetic properties in the quenched BFO–BTO ceramics result in an improved ME response. Figure 5(a) shows the curves of frequency-dependent magnetoelectric coefficient (αME) of the quenched BFO–BTO rectangular ceramic bars at a DC magnetic field of 6000 Oe. During the ME measurement, the applied magnetic field is parallel to the surface of ceramic bars. As shown in Fig. 5(a), αME peaks are observed to be in a frequency range between 70 kHz and 100 kHz for these four samples. The peaks are attributed to the excitation of the mechanical resonance in the BFO–BTO ceramics. In the inset of Fig. 5(a), the impedance spectra of 0.75BFO–0.25BTO ceramics are shown, and the resonance and antiresonance peaks originating from the longitudinal mechanical vibration of rectangular ceramic plates through the transverse piezoelectric response (d31) can be observed. The frequencies for the maximum αME are consistent with the resonance–antiresonance frequencies measured from the impedance spectra. From Fig. 5(a), it is interesting that the maximum ME response is observed in 0.75BFO–0.25BTO ceramics although the 0.7BFO–0.3BTO ceramics have the largest piezoelectric response (Fig. 3(a)) and the strongest ferromagnetism (Fig. 4(a)), the mechanism of which we will discuss later. A strong ME response ∼ 1.3 V/cm·Oe is measured near the resonance frequency of 0.75BFO–0.25BTO ceramic plate. The αME of 0.6BFO–0.4BTO ceramics is negligibly small compared with those with other compositions because the d33 of the composition is small (∼ 19 pC/N, Fig. 3(a)). The plots of dependence of the αME of BFO–BTO ceramics at non-resonance frequency (10 kHz) on DC magnetic field for four different compositions are shown in Fig. 5(b). The αME increases linearly with magnetic field increasing and is typically below 0.025 V/cm·Oe. From Figs. 5(a) and 5(b), we can see that the αME at non-resonance frequency is much smaller than that measured at resonance frequency, reflecting the amplification of the ME response under mechanical resonance conditions.
Figure 5.(a) Frequency-dependent magnetoelectric coupling coefficient
Because the ferromagnetism and piezoelectric effect of BFO-based material originate from different microscopic mechanisms, the ME response is generated not from direct coupling of the two physical effects, but through the mechanical interaction. When αME is measured, the mechanical response is first generated through the magneto-mechanical effect, which, we believe, is the magnetostriction because the sign of αME does not change after the direction of the applies magnetic DC field has been reversed as shown in Fig. 5(b), and then converted into electrical response by the piezoelectric effect.[
Figure 6.Mechanical quality factor and dielectric loss
4. Conclusions
In this work, (1−x)BFO–xBTO ceramics with the compositions near the morphotropic phase boundary is prepared by using a conventional solid-state reaction method and the effect of quenching on the piezoelectric, magnetic, and magnetoelectric properties are investigated. We find that in addition to the piezoelectric response, the quenching of the ceramics from a high temperature can greatly enhance the ferromagnetism of BFO–BTO ceramics. A large piezoelectric response higher than 200 pC/N, which can be sustained at 350 °C, can be achieved in the quenched 0.7BFO–0.3BTO ceramics. Due to the high piezoelectric response and enhanced ferromagnetism of the quenched BFO–BTO ceramics, a greatly enhanced magnetoelectric response up to 1.30 V/cm·Oe is measured in 0.75BFO–0.25BTO ceramics at the mechanical resonance frequency of the sample. The mechanism for the composition dependence of the ME response is also investigated. We show that a high mechanical quality factor and low dielectric loss are desirable to achieve a high ME response, especially under the mechanical resonance conditions.
[1] D I Khomskii. J. Magn. Magn. Mater, 306, 1(2006).
[2] M Fiebig, T Lottermoser, D Frohlich, A V Goltsev, R V Pisarev. Nature, 419, 818(2002).
[3] J Ravez, S C Abrahams, R D Pape. J. Appl. Phys, 65, 3987(1989).
[4] S Sarraute, J Ravez, R VonderMuhll, G Bravic, R S Feigelson, S C Abrahams. Acta. Cryst. B, 52, 72(1996).
[5] J Ravez. J. Phys. III France, 7, 1129(1997).
[6] S W Cheong, M Mostovoy. Nat. Mater, 6, 13(2007).
[7] R Palai, R S Katiyar, H Schm, P Tissot, S J Clark, J Robertson, S A T Redfern, G Catalan, J F Scott. Phys. Rev. B, 77(2008).
[8] G Catalan, J F Scott. Adv. Mater, 21, 2463(2009).
[9] X D Qi, J Dho, R Tomov, M G Blamire, J L MacManus-Driscoll. Appl. Phys. Lett, 86(2005).
[10] J Wang, J B Neaton, H Zheng. Science, 299, 1719(2003).
[11] P Sharma, A Kumar, D Varshney. Solid State Commun, 220, 6(2015).
[12] M Kumara, K L Yadav. Appl. Phys. Lett, 91(2007).
[13] Y H Lee, J M Wu, C H Lai. Appl. Phys. Lett, 88(2006).
[14] J K Kim, S S Kim, W J Kim. Appl. Phys. Lett, 88(2006).
[15] T L Yan, B Chen, G Liu, R P Niu, J Shang, S Gao, W H Xue, J Jin, J R Yang, R W Li. Chin. Phys. B, 26(2017).
[16] N Itoh, T Shimura, W Sakamoto, T Yogo. Ferroelectrics, 356, 19(2007).
[17] M Zhang, X Y Zhang, X W Qi, Y Li, L Bao, H GuY. Ceram. Int, 43(2017).
[18] C Behera, A K Pattanaik. J. Mater. Sci-Mater. El, 089(2019).
[19] G H Ryua, A Hussaina, M H Leea, R A Malik, T K Songa, W J Kim, M H Kim. J. Eur. Ceram. Soc, 18(2018).
[20] X H Du, J H Zheng, U Belegundu, K Uchino. Appl. Phys. Lett, 72, 2421(1998).
[21] Y X Wei, X T Wang, J T Zhu, X L Wang, J J Jia. J. Am. Ceram. Soc, 96, 3163(2013).
[22] T Zhen, Z G Jiang, J G Wu. Dalton Trans, 45(2016).
[23] M H Lee, D J Kim, J S Park, S W Kim, T K Song, M H Kim, W J Kim, D Do. Adv. Mater, 27, 6976(2015).
[24] H He, J T Zhao, Z L Luo, Y J Yang, H Xu, B Hong, L X Wang, R X Wang, C Gao. Chin. Phys. Lett, 33(2016).
[25] K H Zhao, Y H Wang, X L Shi, N Liu, L W Zhang. Chin. Phys. Lett, 32(2015).
[26] W Rao, Y B Wang, Y A Wang, J X Gao, W L Zhou, J Yu. Chin. Phys. Lett, 31(2014).
[27] Y A Wang, Y B Wang, W Rao, J X Gao, W L Zhou, J Yu. Chin. Phys. Lett, 30(2013).
[28] R Gupta, J Shah, S Chaudhary, R K Kotnala. J. Alloys Compd, 638, 115(2015).
[29] L Luo, N Jiang, X Zou, D Shi, T Sun, Q Zheng, C G Xu, K H Lam, D M Lin. Phys. Status Solidi A, 212, 2012(2015).
[30] M Zhang, X Y Zhang, X W Qi, H G Zhu, Y Li, Y H Gu. Ceram. Int, 44(2018).
[31] M I Bichurin, D A Filippov, V M Petrov. Phys. Rev. B, 68(2003).
[32] D A Filippov, M I Bichurin, C W Nan, J M Liu. J. Appl. Phys, 97(2005).
[33] Y M Jia, H S Luo, X Y Zhao, F F Wang. Adv. Mater, 20, 4776(2008).
[34] Q Pan, C Fang, H S Luo, B J Chu. J. Eur. Ceram. Soc, 39, 1057(2019).
[35] Q Pan, B J Chu. J. Appl. Phys, 125(2019).
[36] (1987).
[37] S Unruan, M Unruan, T Monnor, S Priya, R Yimnirun. J. Am. Ceram. Soc, 98, 3291(2015).
[38] M M Kumar, A Srinivas, S V Suryanarayana. J. Appl. Phys, 87, 855(2000).
[39] L Cao, C R Zhou, J W Xu, Q L Li, C L Yuan, G H Chen. Phys. Status Solidi A, 213, 52(2016).
[40] Y Wan, Y Li, Q Li, W Zhou, Q J Zheng, X C Wu, B P Zhu, D M Lin. J. Am. Ceram. Soc, 97, 1809(2014).
[41] M M Kumar, S Srinath, G S Kumar, S V Suryanarayana. J. Appl. Phys, 188, 203(1998).
[42] T H Wang, Y Ding, C S Tu, Y D Yao, K T Wu, T C Lin, H H Yu, C S Ku, H Y Lee. J. Appl. Phys, 109(2011).
[43] R A M Gotardo, D S F Viana, M Olzon-Dionysio, S D Souza, D Garcia, J A Eiras, M F S Alves, L F Cotica, I A Santos, A A Coelho. J. Appl. Phys, 112(2012).
[44] T Fujii, S Jinzenji, Y Asahara. J. Appl. Phys, 64, 5434(1988).
[45] F M Bai, J L Wang, M Wuttig, J F Li, N G Wang, A P Pyatakov, A K Zvezdin, L E Cross, D Viehland. Appl. Phys. Lett, 86(2005).
[46] J Ma, J M Hu, Z Li, C W Nan. Adv. Mater, 23, 1062(2011).
Get Citation
Copy Citation Text
Qi Pan, Bao-Jin Chu. Enhanced ferromagnetism and magnetoelectric response in quenched BiFeO3-based ceramics[J]. Chinese Physics B, 2020, 29(8):
Received: Mar. 28, 2020
Accepted: --
Published Online: Apr. 29, 2021
The Author Email: Bao-Jin Chu (chubj@ustc.edu.cn)