Next Article in Journal
Bond Behavior of Reinforced Concrete Considering Freeze–Thaw Cycles and Corrosion of Stirrups
Previous Article in Journal
In Situ Formation of Laser-Cladded Layer on Thin-Walled Tube of Aluminum Alloy in Underwater Environment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dependence of Modified Butterworth Van-Dyke Model Parameters and Magnetoimpedance on DC Magnetic Field for Magnetoelectric Composites

1
Key Lab of Computer Vision and Intelligent Information System, Chongqing University of Arts and Sciences, Chongqing 402160, China
2
School of Electronic Information and Electrical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
*
Author to whom correspondence should be addressed.
Materials 2021, 14(16), 4730; https://doi.org/10.3390/ma14164730
Submission received: 25 July 2021 / Revised: 12 August 2021 / Accepted: 16 August 2021 / Published: 21 August 2021
(This article belongs to the Section Materials Physics)

Abstract

:
This study investigates the impedance curve of magnetoelectric (ME) composites (i.e., Fe80Si9B11/Pb(Zr0.3Ti0.7)O3 laminate) and extracts the modified Butterworth–Van Dyke (MBVD) model’s parameters at various direct current (DC) bias magnetic fields Hdc. It is interesting to find that both the magnetoimpedance and MBVD model’s parameters of ME composite depend on Hdc, which is primarily attributed to the dependence of FeSiB’s and neighboring PZT’s material properties on Hdc. On one hand, the delta E effect and magnetostriction of FeSiB result in the change in PZT’s dielectric permittivity, leading to the variation in impedance with Hdc. On the other hand, the magnetostriction and mechanical energy dissipation of FeSiB as a function of Hdc result in the field dependences of the MBVD model’s parameters and mechanical quality factor. Furthermore, the influences of piezoelectric and electrode materials properties on the MBVD model’s parameters are analyzed. This study plays a guiding role for ME sensor design and its application.

1. Introduction

Magnetoelectric (ME) materials produce strong ME effects due to the mechanical coupling between magnetostrictive and piezoelectric materials, which has been studied intensively in both theories and experiments [1,2,3,4,5,6]. Such ME effects provide a promising candidate for the highly sensitive DC magnetic field sensor due to its significant variations with external direct current (DC) magnetic field. Dong et al. [7] presented a ME laminate under a constant drive of Hac = 1 Oe, which can reach the limit of detection (LOD) for a DC magnetic field Hdc of 10−4 Oe. Sun et al. [2] reported a novel Nano-Electromechanical System (NEMS) AlN/FeGaB resonator with a high DC magnetic field sensitivity of 280 kHz/Oe and a LOD of 8 × 10−6 Oe. Liu et al. [8] demonstrated a highly sensitive DC magnetic field sensor with a LOD of 2 × 10−5 Oe. Martins et al. [9] showed a Metglas/poly(vinylidene fluoride)/Metglas magnetoelectric laminate with the sensitivity of 30 mV·Oe−1 and resolution of 8 μ Oe for Hdc detection, and its correlation coefficient, linearity and accuracy values reached 0.995, 95.9% and 99.4%, respectively. Yao et al. [10] developed a Metglas/PMNT/Metglas laminate with the LOD of 10 × 10−5 Oe for Hdc detection. Wang et al. [11] also proposed a transformer-type magnetic sensor consisting of soft magnetostrictive alloy FeBSiC/piezoelectric ceramics Pb(Zr,Ti)O3/FeBSiC heterostructure wrapped with both the exciting and sensing coils, which provided the maximum magnetic field sensitivity of 2.12 V/Oe and equivalent magnetic noise of 114 × 10−8 Oe/ HZ (at 1 Hz).
Meanwhile the material property and structure of ME composite have been researched intensively for the magnetic sensor application. As such, the field-dependent characteristics of piezomagnetic coefficient for magnetostrictive material [12], the effect of different magnetostrictive materials on the Hdc sensitivity [13,14,15,16], and the optimum structure of piezoelectric/magnetostrictive composite [17,18,19,20] etc. were reported. Additional to understanding the material property of ME composite, it is essential to study the equivalent electrical parameters of ME composite to further improve the DC magnetic sensor performance. However, few articles have reported and analyzed the Hdc dependence of equivalent electrical parameters based on the modified Butterworth–Van Dyke (MBVD) model of magnetoelectric material, even though this is crucial to guide the conditioning circuit design of the ME sensor. Hence, the exploration of electrical equivalent circuit for ME device in this study facilitates understandings of corresponding electrical resonance behavior, which is beneficial for the design and optimization of impedance matching circuits for ME devices. Additionally, this study is expected to guide the design of the magnetic-field-tuned ultrasonic transducer, which can effectively solve the problem of resonance frequency shift and impedance mismatch of the ultrasonic transducer. It is noted that the MBVD model characterizes the loss mechanisms more accurately compared to the conventional Butterworth–Van Dyke model by considering the effects of additional electrical losses and dielectric losses, which can model the measured results more accurately.
In this paper, we investigate the equivalent circuit of the ME sensor based on the MBVD model of PZT/FeSiB laminated composite. It is noted that Lead zirconate titanate Pb(Zr0.3Ti0.7)O3 (PZT) exhibits the outstanding piezoelectric performance and high mechanical quality factor compared to other piezoelectric materials such as polyvinylidene fluoride (PVDF) and BaTiO3 etc. Meanwhile, the FeSiB (International standard trademark Metglas-2605 S2) possesses a low saturation field and a strong magnetostrictive effect at low magnetic biases Hdc due to its ultrahigh magnetic permeability (i.e., the initial magnetic permeability of 45,000). Correspondingly, magnetostrictive material FeSiB and piezoelectric material PZT are utilized for the ME composite in order to obtain highly magnetic sensing capabilities. In this study, the electrical equivalent circuit parameters of the ME sensor are calculated with the electrical resonance characteristics of measured impedance. Furthermore, the dependences of magnetoimpedance and corresponding MBVD model’s parameters on DC magnetic field are measured and discussed. Such dependences are mainly attributed to the delta E and magnetostrictive effects of FeSiB and correspondingly varied PZT’s dielectric permittivity. Additionally, the effects of the piezoelectric materials and electrode material’s properties on the MBVD model’s parameters are analyzed. The study of electrical equivalent circuit facilitates the understanding of electrical resonance behavior for ME devices, and it plays a crucial role in the design of impedance matching circuits for ME devices. Meanwhile, the controllable impedance and dielectric permittivity of PZT/FeSiB ME composites with DC bias magnetic field have broad potential applications, such as tunable spin filters, storage devices, and magnetic sensor etc.

2. Experiment

The ME sensor consists of PZT/FeSiB laminated composite, where the sizes of magnetostrictive (FeSiB, supplied by Foshan Huaxin Microlite Metal Co., Ltd., Foshan, China) layer and the piezoelectric (PZT, produced by Zibo Yuhai Ceracomp Co., Ltd., Zibo, China) layer are 12 mm × 5 mm × 0.03 mm and 12 mm × 6 mm × 0.8 mm, respectively. First, the PZT plate and FeSiB ribbon are dipped in organic impregnant to clean them. Subsequently, the soft magnetic ribbon FeSiB is bonded with PZT plate by using epoxy glue. Here the West System 105/206 resin/hardener epoxy with a good mechanical property and a low viscosity is utilized to provide strong bondings among layers. The mixture ratio for the epoxy part ‘Resin’ and part ‘hardener’ is specified as 5:1 by the supplier. Then the PZT/FeSiB laminated composite is compacted in a vacuum bag and cured for 12 h at room temperature to further guarantee the strong bonding among layers. The thickness of the epoxy layers is controlled to be less than 5 μm with vacuum bagging techniques, which has been proved to negligibly affect the ME performance, according to previous research [19]. Considering the ease of fabrication and ME performance, the PZT/FeSiB laminated composite is designed to operate in the L–T (i.e., longitudinal–transverse) mode. That is to say, the FeSiB layer is magnetized along the longitudinal direction (i.e., length direction) since the demagnetizing field is much smaller along this direction. Meanwhile the silver electrodes of piezoelectric layer are at its top and bottom surfaces, and the PZT is poled along the transverse direction (i.e., thickness direction).
To measure the impedance of ME composite as a function of the external DC magnetic field Hdc, Hdc is applied along the longitudinal direction of FeSiB layer with a pair of electromagnets driven by a SR830 Lock-In Amplifier. Here the Hdc varies from 0 to 400 Oe, which is calibrated with a Gauss magnetometer (Lake Shore 455 DSP, Columbus, OH, USA). Additionally, when analyzing the dielectric characteristics of the ME sensor, an Impedance Analyzer (4194 A HP Agilent, Santa Clara, CA, USA) is used to measure the magnetoimpedance (Z) of ME composite with the excitation frequency ranged from 125 kHz to 155 kHz.

3. Results and Discussion

Figure 1 shows the impedance Z of the ME sensor as a function of electrical excitation frequency f when the varied DC bias magnetic field is applied along the length direction. As illustrated in the inset of Figure 1, the maximum and minimum impedance as a function of excitation frequency show a strong dependence on DC bias magnetic field.
It is known that the impedance of the ME sensor is defined by [21],
Z = μ 0 μ e f f ε 0 ε e f f
where μ e f f and ε e f f are the effective relative permeability and permittivity, μ 0 and ε 0 are vacuum permeability and permittivity, respectively. The effective relative permittivity ε e f f can be represented as [22].
ε e f f = ε r + d 31 , p 2 E m [ n p z t E p z t ( 1 n p z t ) E m + n p z t E p z t tan ( π f 2 f s ) π f 2 f s 1 ]
where ε r is relative permittivity of piezoelectric material, d31,p is the piezoelectric coefficient, fs is the resonance frequency, E p z t and E m are the Young’s modulus of piezoelectric and magnetostrictive materials, respectively. n p z t and 1 n p z t are the volume fractions of piezoelectric material PZT and magnetostrictive material FeSiB in the ME sensor, respectively.
By applying a DC bias magnetic field to the magnetostrictive material FeSiB, the magnetostriction is produced by FeSiB and transferred to the PZT layer through interfacial coupling. Meanwhile the magnetostrictive stress will also change the Young’s modulus E m   of magnetostrictive material FeSiB and corresponding resonance frequency f s . Correspondingly from the inset of Figure 1, the electromechanical resonance frequency fs of the ME sensor shows a strong dependence on DC bias magnetic field Hdc. Specifically, the resonance frequency fs of the ME sensor is determined by the geometrical dimensions and material parameters (i.e., Young’s modulus and mass density) of both piezomagnetic and piezoelectric materials, and is expressed as [12],
f s = 1 2 l E ¯ ρ ¯
where l is the length of the ME sensor, ρ ¯ and E ¯ are the average density and equivalent Young’s modulus of ME laminate, respectively. For the ME composite, E ¯ and ρ ¯ are determined by [6],
E ¯ = ( 1 n p z t ) E m + n p z t E p z t
ρ ¯ = ( 1 n p z t ) ρ m + n p z t ρ p z t
where ρ p z t and ρ m are the densities of piezoelectric and magnetostrictive materials, respectively.
Here the Young’s modulus of magnetostrictive material FeSiB is given by [23],
E m = σ s e + s m e
where s e , σ ,   s m e are the elastic strain, elastic stress and magnetoelastic strain, respectively. The magnetoelastic strain arises from the magnetic domain reorientation during the varied Hdc [24,25], which results in the change in effective Young’s modulus with Hdc. As a result, the shifts in corresponding resonance frequency (Equation (3)) with Hdc are observed.
According to Equations (1) and (2), the variations in the Young’s modulus E m   of FeSiB and resonance frequency f s with Hdc also lead to the changes in effective relative permittivity and corresponding impedance with Hdc for ME composite. It is noted that the combination of magnetoresistance (MR) and the Maxwell–Wagner effect could also cause the magnetodielectric effect, according to the previous report [26]. However, for our asymmetric PZT/FeSiB laminate, piezoelectric material PZT is covered with the insulating epoxy glue at surface to prevent the current penetrating into the neighboring magnetic ribbon FeSiB. Hence, there is no giant magnetoresistance effect since the sensing current cannot go through the magnetic layers and, correspondingly, no spin dependent scattering phenomenon happens in the ferromagnetic layer. Furthermore, Castel et al. [22] have also reported that the magnetodielectric effect of BaTiO3-Ni laminated composite could reach 10% near the resonance frequency at Hdc = 6 kOe and clarified that the magnetodielectric mechanism of their composites was based on the strain effect instead of the Maxwell—Wagner effect.
It is also interesting to find in Figure 2 that the maximum impedance Zm at the antiresonance frequency (fa) increases to a maximum value at Hdc = 30 Oe, and then decreases with further increasing Hdc, while the minimum impedance Zn at the resonance frequency (fs) varies in the opposite trend. Namely, Zn decreases to a minimum value, and then increases with the increasing Hdc. This is mainly because the capacitance is directly proportional to dielectric permittivity, the minimum capacitance value at fa results in the maximum impedance Zm and the maximum capacitance value at fs leads to the minimum impedance Zn according to Equation (1).
In order to understand the trend of impedance as a function of DC magnetic field, the electromechanical (ME) sensor is characterized with a lumped-parameter equivalent circuit based on the MBVD model, as shown in Figure 3. To characterize the loss from the electrodes, the MBVD model adds two additional loss resistors (i.e., R0 and Rs) to obtain a more accurate model compared with the standard Butterworth–Van Dyke model. It consists of two network branches in parallel, where R0 represents the resistance associated with dielectric losses of the ME sensor, Rs represents the resistance associated with electrical losses of electrode, Rm denotes the resistance associated with mechanical losses, Lm and Cm denote the motional inductance and capacitance, C0 represents the static capacitance formed between top and bottom electrodes of the ME sensor.
The analytical expression of impedance Z ( ω ) and the electrical admittance Y ( ω ) for MBVD model are given by [27,28],
Z ( ω ) = R s + ( R 0 + 1 j ω C 0 ) ( R m + 1 j ω C m + j ω L m ) R 0 + 1 j ω C 0 + R m + 1 j ω C m + j ω L m  
Y ( ω ) = 1 1 1 R m + 1 j ω C m + j ω L m + 1 1 j ω C 0 + R 0 + R s  
The series resonance frequency f s and antiresonance frequency f a can be expressed as [27,28],
f s = 1 2 π L m C m
      f a = 1 2 π C m + C 0 C 0 L m C m = f s C m + C 0 C 0
Using Equations (7), (9) and (10), the model parameter values of C0, R0, Rs, Lm, Cm and Rm are extracted from the measured Z. Table 1 lists all the extracted model parameters. To verify the MBVD model for further design of the conditioning circuit, the simulation of the model is implemented with the electrical simulator Agilent ADS. Figure 4 presents the computed impedance Z and phase based on the extracted model parameters at Hdc = 30 Oe, which shows a good agreement with the measured data.
According to the measured Z with DC bias magnetic field Hdc, the corresponding equivalent circuit parameters (i.e., Cm, Lm, C0, Qs, Rs + Rm, fs and fa) are calculated and analyzed as a function of Hdc, as shown in Figure 5, Figure 6, Figure 7 and Figure 8, respectively.
Specifically, Cm and Lm are given by [27,28],
L m = 1 8 ρ A l t l w 2 ( s 11 E d 31 ) 2 = 1 8 ρ l l t l w ( s 11 E d 31 ) 2
C m = 8 π A l t d 31 2 s 11 E = 8 π l l w l t d 31 2 s 11 E  
where d31, s 11 E , ρ , lw, lt and l are the piezoelectric coefficient, elastic compliance coefficient, density, width, thickness and length of the ME sensor, respectively. A = l l w is the plate area.
It is obvious that the the length l and elastic compliance coefficient s 11 E have strong influences on the Cm and Lm according to Equations (11) and (12). Specifically, due to the stress-strain coupling of interlayers, the magnetostrictive strain produced by FeSiB under varying Hdc results in the change in the length l and elastic compliance coefficient s 11 E for piezoelectric material. As a result, the equivalent electrical parameters Cm and Lm of the ME sensor strongly depend on Hdc and vary in the opposite ways, as illustrated in Figure 5. This is due to the fact that Lm and Cm are proportional to ( s 11 E d 31 ) 2 and d 31 2 s 11 E , respectively.
Furthermore, Cm is proportional to the length l of composite, whereas it is inverse proportional to the elastic compliance coefficient. Since the Young’s modulus is the inverse of elastic compliance coefficient, Cm is determined by both the Young’s modulus E and length l. On one hand due to the stress-strain coupling of the interlayers, the length l increases quickly to a maximum value due to the large piezomagnetic coefficient d33,m of FeSiB and then l reaches the saturation with further increasing Hdc. On the other hand, the Young’s modulus E of the magnetostrictive layer and, corresponding, ME composite decrease initially to a minimum value with the increasing Hdc, and then increases and reaches saturation at large Hdc when Hdc further increases [18]. When Hdc < 60 Oe, the magnetostriction does not attain to saturation, Cm is affected by both Young’s modulus and the length l. However, the effect of length l on Cm is more obvious than that of Young’s modulus due to the large d33,m of FeSiB at the small Hdc, which causes Cm to increase with Hdc and reach a positive peak in low magnetic field Hdc = 60 Oe. When Hdc further increases above 60 Oe, the magnetostriction reaches saturation quickly; however, the Young’s modulus E still varies significantly and plays a dominant role in Cm. Currently, Young’s modulus E and corresponding Cm reach the local minimum values when Hdc further increases to 100 Oe, then Cm gradually increases and reaches saturation with further increasing Hdc due to the variation in E, as shown in Figure 5.
Meanwhile the static capacitance C0 can be also expressed with the following expressions [28]:
C 0 = A ε r ε 0 l t
where ε r and ε 0 are the relative dielectric permittivity and vacuum permittivity of the piezoelectric material, respectively.
When Hdc is applied along the longitudinal direction of the ME sensor, the magnetostrictive material FeSiB expands with the increasing Hdc, which changes dielectric permittivity of piezoelectric material due to the transferred magnetostrictive stress. Correspondingly, C0 varies with the DC magnetic field since C0 is strongly determined by the dielectric permittivity. Yao et al. [10] reported that the dielectric permittivity of Terfenol-D/PZT magnetoelectric composite at the resonant frequency decreased and then increased with increasing dc magnetic field. In this case, C0 varies in a similar trend as function of Hdc since C0 is proportional to the dielectric permittivity. Specifically, it is shown in Figure 6 that the static capacitance C0 of the ME sensor first decreases with the increasing Hdc, and then gradually increases.
The Rm is used to characterize the mechanical loss, which is subject to energy loss in the ME sensor. It can be given as [28],
R m = l t 3 8 A d 31 2 = l t 3 8 l l w d 31 2
R m   is primarily determined by the length l. Correspondingly, the variations in the length l due to the magnetostriction of FeSiB cause the variation in R m with Hdc.
By analyzing the variation in equivalent circuit parameters (i.e., Cm, Lm, C0, Rm etc.) with Hdc, it is found that the varying magnetostrictive strain of FeSiB with Hdc is the main reason for the Hdc dependences of equivalent circuit parameters. Furthermore, it is also noted that these equivalent circuit parameters depend on the actively vibrating area A, since Rm and Lm decrease with the enlarged area, whereas the capacitances C0 and Cm increase with the enlarged area. Such relations are important for designing ME sensors.
Furthermore, the mechanical Quality factor (Q-factor) reflects the capability of the ME sensor to reserve mechanical energy and the corresponding loss of resonant circuit. Q is defined as the ratio of the stored energy to the dissipated energy per cycle during oscillation. According to Lakin’s method, the Q s at series resonance frequency f s and Q p at antiresonance frequency f a can be defined as [27,28],
Q s = 2 π f s L m R s + R m    
Q p = 2 π f p L m R 0 + R m
On one hand, the smaller value of RS is desired to improve the effective mechanical Quality factor Q s of the ME sensor, according to Equation (15). Since RS represents the electrical loss of electrode, the type and quality of the electrode materials directly affect the Q s of the ME sensor. In this case, utilizing the electrode material with high acoustic impedance and low resistivity can reduce RS and improve the effective electromechanical coupling coefficient of the ME sensor. On the other hand, Equation (16) predicts that the high Q p value can be obtained when the ME sensor possesses a low R 0 . Since the dielectric losses R0 of the ME sensor is mainly determined by the dielectric loss of piezoelectric material, it means the smaller dielectric loss results in the larger effective mechanical Quality factor Q p of the ME sensor. Additionally, from Equations (15) and (16), it is found that both Q s at resonance frequency and Q p at antiresonance frequency decrease with the increasing mechanical loss Rm. Hence, the mechanical loss Rm plays a primary role in the energy dissipations of the ME sensor.
Subsequently, the Qs, Qp and corresponding loss as a function of Hdc are experimentally investigated to verify and further understand the above theoretical analysis. It is known that Rm depends on the mechanical energy dissipation t a n δ m e c h of the ME sensor [28], while Qs is inversely proportional to t a n δ m e c h [29]. When the DC magnetic field is applied, the mechanical energy dissipation Rs + Rm of magnetostrictive material FeSiB changes dramatically owing to the non- 180 ° domain wall motions. This results in the varied mechanical quality factor Qs of the ME sensor with Hdc, as shown in Figure 7a. Specifically, the quality factor (Qs) at the series resonance frequency f s decreases from 182 to the minimum value of 160 at the Hdc = 200 Oe and then gradually increases according to the MBVD model. Obviously, the variation in Qs is mainly attributed to the magnetic mechanical loss associated with magnetic domain wall movement and material damping of FeSiB.
Furthermore, the trends of Q p and R0 + Rm as a function of Hdc are similar to that of Q s and Rs + Rm, as shown in Figure 7b. However, the magnitude of antiresonance mechanical quality factor Q p ranges from 234 to 245.6, which is higher than the resonance mechanical quality factor Qs. The differences between Q p and Q s were also reported by in previous literature [30,31]. Finally, the resonance frequency fs and antiresonance frequency fa of ME laminated sensor as a function of varied Hdc are investigated, as shown in Figure 8. Both fs and fa exhibit similar trends with Hdc, which increases with the increasing DC bias field. The obvious shifts of resonance frequency fs and antiresonance frequency fa with Hdc indicate that fs and fa of the ME sensor are adjustable by varying the DC bias magnetic field.

4. Conclusions

In summary, the impedance of the ME sensor (i.e., FeSiB/PZT composite) as a function of DC bias magnetic field is experimentally measured and theoretically analyzed. Meanwhile, the simulation results with the MBVD model of the ME sensor agrees with the measured impedance Z accurately. Specifically, the dependences of extracted MBVD model parameters and the magnetoimpedance effects of the ME sensor on Hdc are observed, which result from the varied magnetostriction and the mechanical energy dissipation of magnetostrictive material FeSiB with Hdc due to the corresponding delta E effect and magnetostrictive effect. Furthermore, the influences of piezoelectric materials property and electrode on the MBVD model parameters are analyzed. The analysis of MBVD model for ME composite is beneficial to the design of analog front-end circuits for the corresponding magnetic sensor, which could further improve the LOD.

Author Contributions

Y.W. designed the project; L.C. developed ME composite and carried out experimental works, L.C. carried out simulation work; all authors analyzed the data and reviewed the manuscript; L.C. and Y.W. wrote the paper; L.C. funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by the National Natural Science Foundation of China (Grant No. 61304255), and the Scientific and Technological Research Program of Chongqing Municipal Education Commission (No. KJZD-K201901301), and the Natural Science Foundation of Chongqing (No.cstc2020jcyj-msxmX0899), and Key projects of technological innovation and application demonstration in Chongqing (Grant No. cstc2018jszx-cyzdX0175).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rupp, T.; Truong, B.D.; Williams, S.; Roundy, S. Magnetoelectric transducer designs for use as wireless power receivers in wearable and implantable applications. Materials 2019, 12, 512. [Google Scholar] [CrossRef] [Green Version]
  2. Leung, C.M.; Li, J.; Viehland, D.; Zhuang, X. A review on applications of magnetoelectric composites: From heterostructural uncooled magnetic sensors, energy harvesters to highly efficient power converters. J. Phys. D Appl. Phys. 2018, 51, 263002. [Google Scholar] [CrossRef]
  3. Li, M.; Matyushov, A.; Dong, C.; Chen, H.; Lin, H.; Nan, T.; Qian, Z.; Rinaldi, M.; Lin, Y.; Sun, N.X. Ultra-sensitive NEMS magnetoelectric sensor for picotesla DC magnetic field detection. Appl. Phys. Lett. 2017, 110, 143510. [Google Scholar] [CrossRef]
  4. Zhang, R.; Zhang, S.; Xu, Y.; Zhou, L.; Liu, F.; Xu, X. Modeling of a magnetoelectric laminate ring using generalized hamilton’s principle. Materials 2019, 12, 1442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Friedrich, R.M.; Zabel, S.; Galka, A.; Lukat, N.; Wagner, J.M.; Kirchhof, C.; Quandt, E.; McCord, J.; Selhuber-Unkel, C.; Siniatchkin, M.; et al. Magnetic particle mapping using magnetoelectric sensors as an imaging modality. Sci. Rep. 2019, 9, 1–11. [Google Scholar] [CrossRef]
  6. Spetzler, B.; Kirchhof, C.; Reermann, J.; Durdaut, P.; Höft, M.; Schmidt, G.; Quandt, E.; Faupel, F. Influenceof the quality factor on the signal to noise ratio of magnetoelectric sensors based on the delta-E effect. Appl. Phys. Lett. 2019, 114, 183504. [Google Scholar] [CrossRef]
  7. Dong, S.X.; Zhai, J.; Xing, Z.; Li, J.F.; Viehland, D. Small dc magnetic field response of magnetoelectric laminate composites. Appl. Phys. Lett. 2006, 88, 82907. [Google Scholar] [CrossRef] [Green Version]
  8. Dong, X.W.; Wang, B.; Wang, K.F.; Wan, J.G.; Liu, J.M. Ultra-sensitive detection of magnetic field and its direction using bilayer PVDF/Metglas laminate. Sens. Actuators A Phys. 2009, 153, 64–68. [Google Scholar] [CrossRef]
  9. Reis, S.; Silva, M.P.; Castro, N.; Correia, V.; Martins, P.; Lasheras, A.; Gutierrez, J.; Barandiarán, J.M.; Rocha, J.G.; Lanceros-Mendez, S. Characterization of metglas/poly(vinylidene fluoride)/metglas magnetoelectric laminates for AC/DC magnetic sensor applications. Mater. Des. 2016, 92, 906–910. [Google Scholar] [CrossRef]
  10. Yao, Y.P.; Hou, Y.; Dong, S.N.; Li, X.G. Giant magnetodielectric effect in Terfenol-D/PZT magnetoelectric laminate composite. J. Appl. Phys. 2011, 110, 14508. [Google Scholar] [CrossRef]
  11. Wang, Y.; Xiao, N.; Xiao, R.; Wen, Y.; Li, P.; Chen, L.; Ji, X.; Han, T. Enhanced dc magnetic field sensitivity for coupled ac magnetic field and stress driven soft magnetic laminate heterostructure. IEEE Sens. J. 2020, 20, 14756–14763. [Google Scholar] [CrossRef]
  12. Bian, L.X.; Wen, Y.M.; Li, P. Dynamic magnetomechanical behavior of TbxDy1–xFey Alloy under small-signal AC drive fields superposed with various bias fields. IEEE Trans. Magn. 2016, 52, 1–5. [Google Scholar] [CrossRef]
  13. Chen, L.; Wang, Y. The effects of the soft magnetic alloys’ material characteristics on resonant magnetoelectric coupling for magnetostrictive/piezoelectric composites. Smart Mater. Struct. 2019, 28, 045003. [Google Scholar] [CrossRef]
  14. Chu, Z.; Shi, H.; Shi, W.; Liu, G.; Wu, J.; Yang, J.; Dong, S. Enhanced resonance magnetoelectric coupling in (1-1) connectivity composites. Adv. Mater. 2017, 29, 1606022. [Google Scholar] [CrossRef] [PubMed]
  15. Lou, G.; Yu, X.; Ban, R. A wide-range DC current sensing method based on disk-type magnetoelectric laminate composite and magnetic concentrator. Sens. Actuator A Phys. 2018, 280, 535–542. [Google Scholar] [CrossRef]
  16. Huong Giang, D.T.; Tam, H.A.; Ngoc Khanh, V.T.; Vinh, N.T.; Tuan, P.A.; Tuan, N.V.; Ngoc, N.T.; Duc, N.H. Magnetoelectric vortex magnetic field sensors based on the metglas/PZT laminates. Sensors 2020, 20, 2810. [Google Scholar] [CrossRef]
  17. Zhang, J.; Kang, Y.; Gao, Y.; Weng, G.J. Experimental investigation of the magnetoelectric effect in NdFeB-driven a-line shape terfenol-D/PZT-5A structures. Materials 2019, 12, 1055. [Google Scholar] [CrossRef] [Green Version]
  18. Li, P.; Wen, Y.M.; Huang, X.; Yang, J.; Wen, J.; Qiu, J.; Zhu, Y.; Yu, M. Wide-bandwidth high-sensitivity magnetoelectric effect of magnetostrictive/piezoelectric composites under adjustable bias voltage. Sens. Actuators A Phys. 2013, 201, 164–171. [Google Scholar] [CrossRef]
  19. Filippov, D.A.; Galichyan, T.A.; Laletin, V.M. Influence of an interlayer bonding on the magnetoelectric effect in the layered magnetostrictive-piezoelectric structure. Appl. Phys. A 2014, 116, 2167–2171. [Google Scholar] [CrossRef]
  20. Chu, Z.; Pourhosseiniasl, M.J.; Dong, S. Review of multi-layered magnetoelectric composite materials and devices applications. J. Phys. D Appl. Phys. 2018, 51, 243001. [Google Scholar] [CrossRef]
  21. Bian, L.X.; Wen, Y.M.; Li, P. Analysis of magneto-mechano-electronic coupling factors in magnetostrictive/piezoelectric laminated composite. Acta Phys. Sin. 2009, 58, 4205–4213. [Google Scholar]
  22. Israel, C.; Petrov, V.M.; Srinivasan, G.; Mathur, N.D. Magnetically tuned mechanical resonances in magnetoelectric multilayer capacitors. Appl. Phys. Lett. 2009, 95, 072505. [Google Scholar] [CrossRef]
  23. Cullity, B.D.; Graham, C.D. Introduction to Magnetic Materials, 2nd ed.; John Wiley & Sons: Hoboken, NJ, USA, 2009; p. 270. [Google Scholar]
  24. Clark, A.E.; Savage, H.T. Giant magnetically induced changes in the elastic moduli in Tb.3Dy.7Fe2. IEEE Trans. Sonics Ultrason. 1975, 22, 50–51. [Google Scholar] [CrossRef]
  25. Chen, Z.; Su, Y. The influence of low-level pre-stressing on resonant magnetoelectric coupling in Terfenol-D/PZT/Terfenol-D laminated composite structure. J. Appl. Phys. 2014, 115, 193906. [Google Scholar] [CrossRef]
  26. Catalan, G. Magnetocapacitance without magnetoelectric coupling. Appl. Phys. Lett. 2006, 88, 102902. [Google Scholar] [CrossRef]
  27. Luan, G.D.; Zhang, J.D.; Wang, R.Q. Piezoelectric Transducer and Transducer Array; Peking University Press: Beijing, China, 2005; p. 114. [Google Scholar]
  28. Buttry, D.A.; Ward, M.D. Measurement of interfacial processes at electrode surfaces with the electrochemical quartz crystal microbalance. Chem. Rev. 1992, 92, 1355–1379. [Google Scholar] [CrossRef]
  29. Yang, F.; Wen, Y.M.; Li, P.; Zheng, M.; Bian, L.X. Resonant magnetoelectric response of magnetostrictive/piezoelectric laminate composite inconsideration of losses. Sens. Actuators A Phys. 2008, 141, 129–135. [Google Scholar] [CrossRef]
  30. Zhuang, Y.; Ural, S.O.; Rajapurkar, A.; Tuncdemir, S.; Amin, A.; Uchino, K. Derivation of piezoelectric losses from admittance spectra. Jpn. J. Appl. Phys. 2009, 48, 041401. [Google Scholar] [CrossRef]
  31. Wang, Y.; Gray, D.; Berry, D.; Gao, J.; Li, M.; Li, J.; Viehland, D. An extremely low equivalent magnetic noise magnetoelectric sensor. Adv. Mater. 2011, 23, 4111–4114. [Google Scholar] [CrossRef]
Figure 1. Impedance curve of the ME sensor at various bias DC magnetic fields, and the insets show enlarged details around the maximum and minimum impedances. The maximum relative standard deviations (RSD, i.e., standard deviation/mean × 100%) of impedance with multiple measurements is 0.27%.
Figure 1. Impedance curve of the ME sensor at various bias DC magnetic fields, and the insets show enlarged details around the maximum and minimum impedances. The maximum relative standard deviations (RSD, i.e., standard deviation/mean × 100%) of impedance with multiple measurements is 0.27%.
Materials 14 04730 g001
Figure 2. The maximum impedance Zm and minimum impedance Zn as a function of DC magnetic field. The maximum RSDs of maximum impedance and minimum impedance with multiple measurements are 0.24% and 0.25%, respectively.
Figure 2. The maximum impedance Zm and minimum impedance Zn as a function of DC magnetic field. The maximum RSDs of maximum impedance and minimum impedance with multiple measurements are 0.24% and 0.25%, respectively.
Materials 14 04730 g002
Figure 3. The modified Butterworth–Van Dyke (MBVD) circuit for the ME sensor.
Figure 3. The modified Butterworth–Van Dyke (MBVD) circuit for the ME sensor.
Materials 14 04730 g003
Figure 4. The (a) impedance and (b) phase angle of the ME sensor as a function of the electrical excitation frequencies ranged from 125 kHz to 155 kHz at Hdc = 30 Oe. The maximum RSDs of measured impedances and phase angles are 0.21% and 0.23%, respectively.
Figure 4. The (a) impedance and (b) phase angle of the ME sensor as a function of the electrical excitation frequencies ranged from 125 kHz to 155 kHz at Hdc = 30 Oe. The maximum RSDs of measured impedances and phase angles are 0.21% and 0.23%, respectively.
Materials 14 04730 g004
Figure 5. The Cm and Lm of the ME sensor as a function of DC magnetic field. Here, the maximum RSDs of Cm and Lm with multiple measurements are 0.23% and 0.25%, respectively.
Figure 5. The Cm and Lm of the ME sensor as a function of DC magnetic field. Here, the maximum RSDs of Cm and Lm with multiple measurements are 0.23% and 0.25%, respectively.
Materials 14 04730 g005
Figure 6. C0 of the ME sensor as function of DC magnetic field. The maximum RSD of C0 with multiple measurements is 0.25%.
Figure 6. C0 of the ME sensor as function of DC magnetic field. The maximum RSD of C0 with multiple measurements is 0.25%.
Materials 14 04730 g006
Figure 7. (a) The Qs and Rs + Rm of the ME sensor as a function of DC magnetic field; (b) The Qp and R0 + Rm of the ME sensor as a function of DC magnetic field. The maximum RSDs of Qs and Qp with multiple measurements are 0.22% and 0.21%, respectively. The maximum RSDs of Rs + Rm and R0 + Rm with multiple measurements are 0.24% and 0.26%, respectively.
Figure 7. (a) The Qs and Rs + Rm of the ME sensor as a function of DC magnetic field; (b) The Qp and R0 + Rm of the ME sensor as a function of DC magnetic field. The maximum RSDs of Qs and Qp with multiple measurements are 0.22% and 0.21%, respectively. The maximum RSDs of Rs + Rm and R0 + Rm with multiple measurements are 0.24% and 0.26%, respectively.
Materials 14 04730 g007
Figure 8. Resonance frequency fs and antiresonance frequency fa of ME resonator as a function of DC magnetic field. The maximum RSDs of fs and fa with multiple measurements are 0.23% and 0.27%, respectively.
Figure 8. Resonance frequency fs and antiresonance frequency fa of ME resonator as a function of DC magnetic field. The maximum RSDs of fs and fa with multiple measurements are 0.23% and 0.27%, respectively.
Materials 14 04730 g008
Table 1. Parameters of the equivalent circuit model for ME resonator at Hdc of 30 Oe.
Table 1. Parameters of the equivalent circuit model for ME resonator at Hdc of 30 Oe.
Model ParametersRmLmCmR0RsC0
values88.3 Ω30.2 mH43.8 pF20.8 Ω60 Ω1.08 nF
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, L.; Wang, Y. Dependence of Modified Butterworth Van-Dyke Model Parameters and Magnetoimpedance on DC Magnetic Field for Magnetoelectric Composites. Materials 2021, 14, 4730. https://doi.org/10.3390/ma14164730

AMA Style

Chen L, Wang Y. Dependence of Modified Butterworth Van-Dyke Model Parameters and Magnetoimpedance on DC Magnetic Field for Magnetoelectric Composites. Materials. 2021; 14(16):4730. https://doi.org/10.3390/ma14164730

Chicago/Turabian Style

Chen, Lei, and Yao Wang. 2021. "Dependence of Modified Butterworth Van-Dyke Model Parameters and Magnetoimpedance on DC Magnetic Field for Magnetoelectric Composites" Materials 14, no. 16: 4730. https://doi.org/10.3390/ma14164730

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