ArticlePDF Available

Review of Magnetoelectric Sensors

Authors:

Abstract and Figures

Multiferroic magnetoelectric (ME) materials with the capability of coupling magnetization and electric polarization have been providing diverse routes towards functional devices and thus attracting ever-increasing attention. The typical device applications include sensors, energy harvesters, magnetoelectric random access memory, tunable microwave devices and ME antenna etc. Among those application scenarios, ME sensors are specifically focused in this review article. We begin with an introduction of materials development and then recent advances in ME sensors are overviewed. Engineering applications of ME sensors were followed and typical scenarios are presented. Finally, several remaining challenges and future directions from the perspective of sensor designs and real applications are included.
Content may be subject to copyright.
Actuators 2021, 10, 109. https://doi.org/10.3390/act10060109 www.mdpi.com/journal/actuators
Review
Review of Magnetoelectric Sensors
Junqi Gao 1,2,3, Zekun Jiang 1,2,3, Shuangjie Zhang 1,2,3, Zhineng Mao 1,2,3, Ying Shen 1,2,3,* and Zhaoqiang Chu 1,2,3,*
1 Acoustic Science and Technology Laboratory, Harbin Engineering University, Harbin 150001, China;
gaojunqi@hrbeu.edu.cn (J.G.); jjk95@hrbeu.edu.cn (Z.J.); zhangshuangjie@hrbeu.edu.cn (S.Z.);
maozhineng@hrbeu.edu.cn (Z.M.)
2 Ministry of Industry and Information Technology, Key Laboratory of Marine Information Acquisition and
Security, Harbin Engineering University, Harbin 150001, China
3 College of Underwater Acoustic Engineering, Harbin Engineering University, Harbin 150001, China
* Correspondence: shenying@hrbeu.edu.cn (Y.S.); zhaoqiangchu@hrbeu.edu.cn (Z.C.)
Abstract: Multiferroic magnetoelectric (ME) materials with the capability of coupling magnetization
and electric polarization have been providing diverse routes towards functional devices and thus
attracting ever-increasing attention. The typical device applications include sensors, energy harvest-
ers, magnetoelectric random access memory, tunable microwave devices and ME antenna etc.
Among those application scenarios, ME sensors are specifically focused in this review article. We
begin with an introduction of materials development and then recent advances in ME sensors are
overviewed. Engineering applications of ME sensors were followed and typical scenarios are pre-
sented. Finally, several remaining challenges and future directions from the perspective of sensor
designs and real applications are included.
Keywords: multiferroic; magnetoelectric; sensors; object detection; magnetic localization; current
sensing; biological magnetic measurement; non-destructive testing; displacement sensing
1. Introduction
Multiferroic materials have been recently attracting ever-increasing attention be-
cause of the capability of coupling at least two ferric orders, i.e., ferroelectricity, ferromag-
netism, or ferroelasticity, and the vast potential for multifunctional devices applications
[1–5]. A control of polarization P by external magnetic field H (direct ME (DME) effect) or
a manipulation of magnetization M by an electric field E (converse ME (CME) effect) can
be realized in multiferroic magnetoelectric (ME) materials [6]. Compared with single-
phase ME material, ME heterostructures and ME laminates perform greatly enhanced
coupling capability, which is generally characterized by ME coefficient  [7–9]. After
a development of nearly half a century, tremendous progress regarding ME composites
and related device applications has been reported [1–3,6,10–19].
In this article, the focus is placed on magnetoelectric sensors and the corresponding
engineering applications. After an overview of materials fundaments, we present current
advances in ME sensors including DC, low-frequency and resonant magnetic field sens-
ing. Then we summarize typical engineering applications of ME sensors including object
detection and localization, speed and displacement sensing, current sensing and non-de-
structive testing, stress and strain sensing, and biological magnetic measuring. We will
also discuss some remaining questions of ME sensors and their engineering applications
at the end of the article.
2. Materials for ME Sensors
The ME effect was first experimentally demonstrated in single-phase multiferroic
material Cr2O3 in 1961 [20,21]. After that, diverse studies all over the globe were con-
ducted to further enhance the coupling capability of ferroelectric and magnetic orderings
Citation: Gao, J.; Jiang, Z.; Zhang, S.;
Mao, Z.; Shen, Y.; Chu, Z. Review of
Magnetoelectric Sensors. Actuators
2021, 10, 109. https://doi.org/10.3390/
act10060109
Academic Editor: Shuxiang Dong
Received: 26 April 2021
Accepted: 19 May 2021
Published: 24 May 2021
Publisher’s Note: MDPI stays neu-
tral with regard to jurisdictional
claims in published maps and insti-
tutional affiliations.
Copyright: © 2021 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and con-
ditions of the Creative Commons At-
tribution (CC BY) license (http://cre-
ativecommons.org/licenses/by/4.0/).
Actuators 2021, 10, 109 2 of 23
in a single-phase material system [20,22], but the low Curie temperature and the weak ME
coupling capability in single-phase ME materials, such as BiFeO3, BiMnO3 and LuFe2O4,
greatly limited their applications [1,23,24]. The proposal of a product effect in composite
ME materials by combining the piezomagnetic and piezoelectric effects of ferromagnetic
and ferroelectric materials then provided new routes towards improved ME coupling per-
formance. Early in 1986, Pantinakis et al. proposed 2-2 type ME composites based on the
aforementioned product effect [25] and giant ME coefficients were gradually realized in
laminated ME composites starting from the beginning of 21st century [1,6,10]. Compared
with single-phase or 0-3 typed ME materials, 2-2 typed ME composites, such as a bulk ME
laminates with piezoelectric phase (Pb(Zr,Ti)O3(PZT), Pb(Mg,Nb)O3-PbTiO3 (PMN-PT))
embedded in piezomagnetic materials (FeCoSiB, FeBSiC Terfenol-D, Ni or Fe-Ga) [8] and
a FeGaB/AlN thin-film ME heterostructure [26], exhibited enhanced ME coupling perfor-
mance benefitting from the removal of the leakage current and the improvement of the
interfacial strain transfer. At this section, we will first review materials advances in ME
sensors since 2002.
2.1. Bulk ME Laminates
It is highly desirable to design new connectivity structures for circumventing the lim-
itation of leakage current that occurs in 0-3 typed ME composites. Back in 2002, Ryu et al.
developed a laminated Terfenol-D/PZT/Terfenol-D ME composite (Figure 1a) with 2-2
type connectivity to solve the leakage current problem in 0-3 type ME composites, and the
obtained ME coupling coefficient at non-resonance frequency reached as high as 5
V/cm·Oe [27]. This was a significant event in the development of ME laminates and vari-
ous kinds of laminated structures were proposed afterwards [10,27]. For example. Dong
et al. reported 2-2 type ME laminates consisting of Terfenol-D ferrite and PMN-PT piezo-
electric crystal. These ME composites work with L-T mode and display relatively low ME
coefficients of 2.2 V/cm·Oe at non-resonance frequency [28]. In a bid to further improve
the ME voltage coefficient, Dong et al. in 2005 first proposed a push-pull mode that in-
creased the distance between electrodes and decreased the static capacitance of ME lami-
nates from nF to pF scale [29,30]. In such 2-2 type ME composites, the piezoelectric core
was symmetrically poled along its longitudinal direction and rgw d33 piezoelectric con-
stant of a piezoelectric material could be utilized. A giant ME voltage coefficient of 1.6
V/Oe at non-resonant frequencies was observed experimentally [30]. One year later, Dong
et al. further developed a multi-push-pull mode in 2-1 ME composites. The schematic
structure configuration and operation mode of such a 2-1 ME composite is presented in
Figure 1c. It consisted of a piezo-fiber layer laminated between FeBSiC alloys. For the first
time, the non-resonant ME coefficient at 1 Hz reached 22 V/cm·Oe, making such a struc-
ture especially suitable for low-frequency and passive magnetic sensing [31–35], but it
should be noted here that the mechanical quality factor for such a 2-1 type ME composites
is normally less than 100, so ultra-high resonant ME coefficients cannot be realized in this
case [29].
Another way to address the difficulty of fully polarizing the piezoelectric phase in 0-
3 type ME composites is replacing the particle phase with a 1-D piezoelectric fiber (form-
ing 1-3 typed connectivity). For example, in 2005 Nan et al. reported a 1-3 type ME com-
posite with ZT rod arrays embedded in a Terfenol-D medium via a dice-and-fill technique.
The non-resonant ME coupling coefficient reached 6.2 V/cm·Oe [36], which represented
great progress for ME composites. Two years later, Ma et al. simplified this 1-3 type ME
structure by just embedding one single PZT rod in a Terfenol-D/epoxy mixture [37]. The
single period element of the 1-3 ME composites is shown in Figure 1b. Although the non-
resonant ME coupling coefficient decreased by almost one order of amplitude, this simple
structure, low-cost fabrication process and sub-millimeter size made it attractive for mi-
cro-ME array applications [37].
Actuators 2021, 10, 109 3 of 23
Figure 1. (a) Schematic structure (top) and photograph (bottom) of ME laminate composites using Terfenol-D and PZT
disks [27]. (b) 3D and crosss ectional schematic illustration of the single period of 1-3-type ME structure [37]. (c) Illustration
of the FeBSiC/piezofiber laminate configuration working on multi-push-pull mode [29,30]. (d) The schematic view for 1-1
laminated ME composite and a-(ii) the prototype snapshot of the 1-1 typed ME sample [8].
In 2017, Chu et al. reported a 1-1 type ME composites, which consisted of a [011]-
oriented Pb(Mg,Nb)O3-PbZrO3-PbTiO3 (PMN-PZT) single crystal fiber and laser-treated
amorphous alloy Metglas. The 1-1 type ME composite featured athe one-dimensional con-
figuration as shown in Figure 1d [8]. The laser treatment could decrease magnetic hyste-
resis loss of Metglas and thereby enhance the Q value of the ME resonator. In addition,
the fiber configuration effectively utilized the magnetic flux concentration effect occurring
in Metglas layers. More importantly, this 1-D configuration favored the longitudinal vi-
bration mode of ME laminates. A ME coupling coefficient of ~7000 V/cm · Oe, that was
nearly seven times higher than the best result published previously, was finally realized,
opening a door to develop new ME devices, e.g., resonant magnetic receivers in particular
[8]. In addition, a high ME coefficient of 29.3 V/cm·Oe at non-resonant frequency was also
achieved for our 1-1 type composites. Note, only one single crystal was consumed in this
case, while previous 2-1 type composites normally took five crystals. In 2020, the resonant
ME coefficient of 1-1 type ME composites was further enhanced to 12,500 V/cm·Oe by
using a hard piezo-crystal Mn-PMN-PZT [9]. A summary of the field coupling coefficient
of different ME laminates, i.e., 0-3, 2-2, 2-2.1-1 ME laminates, is given in Table 1.
With respect to ceramic-based thin film multiferroic laminates, Ryu et al. recently
developed a Pb(Zr,Ti)O3 film deposited on piezomagnetic materials, e.g., Ni and Metglas.
The crystallization of PZT film was implemented by laser annealing, which was able to
keep the piezomagnetic layer free from property degradation [38–41]. Readers can get
access to more detailed information concerning film-based ME composites in other review
papers [3,6,10].
Actuators 2021, 10, 109 4 of 23
Table 1. Some ME laminates and their ME coupling performances.
Composition Year Connectivity Working Mode


(V/cm·Oe)


(V/cm·Oe)
Terfenol-D/PZT [37] 2007 3-1 L-L 0.5 18.2
NiFe2O4/PZT [42] 2001 2-2 L-T 1.5 /
Terfenol-D/PZT [27] 2002 2-2 L-T 5 /
Metglas/PVDF [43] 2006 2-2 L-T 7.2 310
Metglas/P(VDF-TrFE) [44] 2011 2-2 L-L 17.7 383
Lanthanum gallium tantalite/
permendur [45] 2012 2-2 / 2.3 720
FeCoSiB/(Pt)/AlN in vacuum [46] 2013 2-2 L-T / 20,000
FeCoSiB/(Pt)/AlN [47] 2016 2-2 L-T / 5000
Metglas/LiNbO3 [48] 2018 2-2 L-T 1.9 1704
FeBSiC/PZT [30] 2006 2-1 L-L 22 500
Metglas/PMN-PT [31] 2011 2-1 L-L 45 1100
Metglas/PMN-PT without laser
treatment [8] 2017 1-1 L-T 29.3 5500
Metglas/PMN-PT with laser
treatment [8] 2017 1-1 L-T 22.9 7000
Metglas/Mn-PMN-PZT with laser
treatment [9] 2020 1-1 L-T 23.6 12,500
Note: Connectivity. We use different number to represent the connectivity of each individual phase. For example, 1-3 type
composite means one-phase fiber (denoted by 1) was embedded in the matrix of another phase (denoted by 3); 2-2 type
composite means laminated structure (each phase has a plane configuration denoted by 2); 2-1 type composite means one-
phase fiber was laminated with another phase plate; 1-1 type means both phases are in the form of fiber configuration.
Working mode. L-L, L-T means longitudinal vibrations with longitudinal magnetization and transverse polarization(L-L)
or transverse magnetization and transverse polarization (L-T).
2.2. MEMS and NEMS ME Laminates
In a bid to obtain miniaturized ME devices with enhanced ME coupling capability,
micro-electro-mechanical systems (MEMS) fabrication technology is a promising ap-
proach benefiting from the strong interfacial bonding force and the fine control over the
material composition. Greve et al. developed a thin film MEMS composite consisting of
AlN and amorphous Fe90Co78Si12B10 [49]. AlN is an ideal piezoelectric material compatible
with MEMS techniques, and amorphous soft magnetic alloy is a good candidate for the
piezomagnetic phase because of its high piezomagnetic properties. As shown in Figures
2a,b, two kinds of deposition flow could be used for MEMS ME composites. Conventional
process flow involves the deposition of a high temperature constituent (AlN). In Figure
2a, a reverse flow was then proposed, where FeCoSiB was deposited as the first layer on
the smooth wafer surface and AlN, including with the Pt seed layer, was deposited on top
of it without any substrate heating [47]. A giant ME coupling coefficient of 5000 V/cm·Oe
was measured in this case [47]. In Figure 2b, depositing the magnetostrictive layer and the
piezoelectric layer on two sides of a silicon substrate separately is another way to obtain
good MEMS ME films [50]. With respect to NEMS ME films, Sun’s group in Northeastern
University has contributed lots of works in this field [26,51,52]. As shown in Figures 2c,d,
the typical material is AlN and FeGaB film. As a ME resonator, both laterally-vibrating
(Figure 2c) or vertically-vibrating (Figure 2d) mode can be realized at different frequency
bands. Recently, a NEMS ME resonator has been successfully utilized for mechanical an-
tennas with miniaturized size compared with traditional antennas driven by RF current
[53].
Actuators 2021, 10, 109 5 of 23
Figure 2. Sketch of ME MEMS cantilever with the functional layer deposited on one side (a) [47] and two side (b) [50] of
silicon substrate. (c) Scanning electron microscopy (SEM) images of the ME nano plate resonator. (d) Scanning electron
microscopy (SEM) images of the fabricated ME thin-film bulk acoustic wave resonators. The red and blue areas show the
suspended circular plate and AlN anchors. The yellow area presents the electrode [53].
3. Advances in ME Sensors
The giant ME coupling in ME composites provides the chances to be implemented as
diverse functional devices, such as sensors, energy harvesters, magnetoelectric random
access memories, tunable microwave devices and ME antennas, etc. Among those appli-
cation scenarios, advances in ME sensors will be reviewed in this section.
To assess the performance of a general magnetic sensor, several critical parameters
should be considered, i.e., limit of detection (LoD), sensitivity, working temperature, dy-
namic range, power consumption, size and the cost, but one should note LoD and sensi-
tivity should be given a high priority when taking the research stage of ME sensors into
consideration. With respect to the LoD of ME sensors, the ME coupling coefficient and the
voltage noise level should be considered equally. Table 1 summarizes the ME coefficients
of typical ME composites. The total noise level comes from both internal and external
noise sources. The internal noise is dominated by the dielectric loss  and the leakage
resistance, which can be written as follows [32,33]:
=
+
=
 +
()
,
(1)
where k is the Boltzmann constant (1.38 × 10−23 J K−1), T is the temperature in Kelvin, Cp is
the static capacitance, tan δ is the dielectric loss, f is the frequency in Hz and R is the DC
resistance of the ME sensor. The total noise density has a 1/f spectrum and makes the
magnetic field detection at low frequency much more difficult. On the other hand, ME
sensors are susceptible to external environment variations, e.g., temperature fluctuation
and base vibration, which typically occurs in low frequency as well [7,54]. We will discuss
the current advances in ME sensors focusing on the improvement of LoD in the following
sections.
Actuators 2021, 10, 109 6 of 23
3.1. Low-Frequency Magnetic Sensor
In 2011, Wang et al. reported the realization of an extremely low limit of detection
through a combination of giant ME coupling in 2-1 type ME composites and a reduction
in each noise source. Giant ME coupling was achieved by optimizing the stress transfer in
multi-push-pull mode, the thickness ratio of Metglas to piezofiber, and the ID electrodes
distribution on Kapton (Figure 3a). Experimental results showed that an extremely low
equivalent magnetic noise of 5.1 pT/√Hz at 1 Hz was obtained (Figure 3b) [33].
Figure 3. (a) Schematic diagram and the protype photo of 2-1 type ME composite working on multi-push-pull mode. (b)
Measured and estimated equivalent magnetic noise of the proposed sensor unit [33].
The problem in 2-1 type ME composites based on multi-push-pull working mode is
the difficulty to fully polarize the piezoelectric phase and the capacitance in this configu-
ration is usually small. In 2012, Li et al. further pointed out that the equivalent magnetic
noise could be reduced by a factor of √N through stacking some number N of ME sensor
units in parallel [32]. From the perspective of reducing the total noise level , connecting
N ME sensor units in series could be also effective to increase the detection capability. For
example, Fang et al. reported a 2-1 ME sensor based on multi-L-T mode, of which the
schematic is shown in Figures 4a,b [55]. In this case, the ME charge coefficient could be
kept at a high level while the static capacitance and the leakage current could be decreased
remarkably by increasing the number (N) of piezoelectric crystal. As a result, the meas-
ured equivalent magnetic noise (EMN) of the Metglas/Mn-PMNT composite was as low
as 0.87 pT/√Hz at 30 Hz for N = 7, which was 1.8 times lower than that for N = 1 (see
Figures 4c,d) [55].
In 2011, frequency conversion technology (FCT) was proposed to circumvent the
large 1/f noise for active ME sensors [56–60]. Quasi-static or extremely-low frequency
magnetic fields can be effectively detected in this case. For example, Chu et al. realized a
limit of detection of 33 pT/√Hz at 0.1 Hz by using amplitude modulation method com-
bined with FCT in 1-1 type magnetoelectric composites [61]. During the measurement, a
carrier signal and a modulation signal were both applied to the ME sensor.
Actuators 2021, 10, 109 7 of 23
Figure 4. 3D structure of Metglas/Mn-PMNT ME composite (a) and its cross-sectional diagram (b); (c) The EMN over the
frequency range of 8 Hz < f < 100 Hz. (d)The EMN and of different Metglas/Mn-PMNT sensors at 30 Hz [55].
Figures 5a,b demonstrates the fundamental modulation phenomenon and the block
diagram of the correlation detection scheme with respect to an amplitude modulation sig-
nal SMod (t). The output voltage waveform was observed by a mixed signal oscilloscope.
The ME sensor was driven by 100 Hz carrier signal and the modulation frequency is 10
Hz. Once the low-frequency modulation field HAC with an intensity of 10−6 T was applied,
a clear amplitude modulation (envelope) signal was generated due to the intrinsic fre-
quency mixing characteristic in ME sensors, as shown in Figure 5a(ii).
In order to test the limit of detection by using this amplitude modulation method, the
time constant decreased to 10 ms and the demodulated signal from time domain wave-
form via a lock-in amplifier was analyzed. Figure 5c shows the measured output voltage
in response to an applied 100 mHz HAC varying from 0.1 to 10 nT. Clearly, a standard
linear-response to HAC within this range was obtained as given in the inset in Figure 5c.
Accordingly, the limit of resolution (LOR) of the ME sensor based on this amplitude mod-
ulation method was determined to be as low as 100 pT. To confirm this LOR, Figure 5d
further verified it by measurement. Considering an equivalent noise bandwidth
(ENBW)of 7.8 Hz corresponding to the given measurement system, the calculated LoD
was then calculated as 33 pT/√Hz at 0.1 Hz.
Actuators 2021, 10, 109 8 of 23
Figure 5. (a) The demonstration of fundamental modulation and frequency mixing phenomenon in ME sensors; (b) A
block diagram of the amplitude demodulation method with respect to amplitude modulation signal SMod (t). (c) The meas-
ured output waveform in response to an applied weak AC magnetic field at 100 mHz. (d) A linear-response to varying
HAC at 100 mHz with a step of 0.1 nT [61].
3.2. Resonant-Frequency Magnetic Sensor
ME laminates can be viewed as resonators from the perspective of mechanics and
resonant phenomenon is also able to enhance the ME coupling coefficient and thus to im-
prove the detection ability [10]. In this regard, ME sensors could be highly competitive
over other magnetic field sensors, e.g., fluxgate sensor and optical pump magnetometer.
Using a 2-2 ME composite, Dong et al. reported an enhanced LoR of 1.2 pT early in 2005
(see Figure 6a) [29]. As for MEMS ME magnetic sensor, Yarar et al. developed a low tem-
perature deposition route of very high quality AlN film, allowing the reversal process
flow as talked in Section 2.2. Correspondingly, the LoD was enhanced by almost an order
of magnitude approaching 400 fT/Hz1/2 at the electromechanical resonance, as shown in
Figure 6b [47]. Based on the giant resonance ME coupling coefficient in 1-1 type ME lam-
inate, a superhigh resonant magnetic-field sensitivity close to be 135 fT (see Figure 6c) was
further obtained by Chu et al. [8], which indicates great potential for 1-1 type ME compo-
sites in the field of eddy current sensing, space magnetic sensing and active magnetic lo-
calizing [8,62]. In 2018 Turutin et al. reported a new ME composite consisting of the y +
140° cut congruent lithium niobate piezoelectric plates with an antiparallel polarized
“head-to-head” bidomain structure and magnetostrictive material Metglas [48]. Based on
this 2-2 ME bimorph, the equivalent magnetic noise spectral density was only 92 fT/Hz1/2
and the directly measured resolution was found to be 200 fT at a bending resonance fre-
quency of 6862 Hz (see Figure 6d), but one should note that the bandwidth of resonant
ME sensors is normally below 1 kHz due to the high mechanical quality factor, which is a
major limitation facing practical engineering applications [8,48,63]. It should however be
noted that resonant ME sensors are greatly limited by the narrow bandwidth and specifi-
cally suited applications need to be considered.
Actuators 2021, 10, 109 9 of 23
Figure 6. (a) Magnetic field detection limit measurements at frequencies of f = 1 Hz and f = 77.5 kHz (resonance condition),
respectively [29]; (b)The measurement of LOD for MEMS ME sensor [47], (c) for 1-1 typed ME sensor [8] and (d) for a 2-2
ME bimorph [48].
3.3. DC Magnetic Sensor
DC or quasi-static magnetic sensors are promising for magnetic anomaly detection
uses, such as geomagnetic navigation, metal detection and magnetic medical diagnosis,
etc. Early in 2011, Gao et al. demonstrated the excellent detection ability for DC field using
2-1 ME composite [31]. As shown in Figures 7a,b, the magnetic resolution was found to
be 4 nT and 1 nT when driving the composite at non-resonant frequency and resonance
frequency, respectively [31]. In 2013, Nan et al. reported a self-biased 215 MHz magneto-
electric NEMS resonator consisting of an AlN/(FeGaB/Al2O3) multilayered heterostructure
(Figure 7c), for ultra-sensitive DC magnetic field detection [51]. A ultra-sensitive detection
level starting from 300 picoTesla was obtained experimentally (Figure 7d) [51]. The RF
NEMS magnetoelectric sensor is compact, power efficient and readily integrated with
CMOS technology, however, the measurement of the resonance frequency and the admit-
tance spectrum is not technologically convenient. Li et al. then further proposed to moni-
tor the reflected output voltage from the ME resonator directly [26]. The optimized detec-
tion sensitivity was determined as 2.8 Hz/nT for AlN/FeGaB resonator. An ultra-high fre-
quency (UHF) lock-in amplifier and a directional coupler were used to apply and test the
RF signal of this resonator. And the final limit of detection was measured to be around 0.8
nT.
Actuators 2021, 10, 109 10 of 23
Figure 7. The measurement of LoD for Metglas/PMN–PT ME laminate at (a) f = 10 kHz and (b) resonance frequency of
27.778 kHz [31]. (c) Schematic representation and (d) the measurement of LoD for NMES AlN/(FeGaB/Al2O3) multilayered
heterostructure [51]; (e) Schematic representation of the conventional flux gate senor and the proposed ME flux gate sensor
[64]; (f) The measured results for DC magnetic field resolution [64].
Using the nonlinear resonance magnetoelectric effect in ME composites, Burdin et al.
fabricated a planar langatate-Metglas structure and employed the third harmonics of the
output signal to measure the DC magnetic field as low as 10 nT [65]. In addition, a broad
dynamic range from ~10 nT to about 0.4 mT was also successfully obtained using the non-
linear ME effect [66]. More recently, Chu et al. proposed a shuttle-shaped, non-biased
magnetoelectric flux gate sensor (MEFGS) for DC magnetic field sensing enlightened by
the design of conventional flux gate sensor [64]. Figure 7e shows both the schematic of
typical flux gate senor and the proposed magnetoelectric flux gate sensor. The flux gate
sensor based on Faraday’s Law of Induction is composed of a racetrack type magnetic
core surrounded by an excitation (first) coil and a detection (second) coil. With respect to
MEFGS, a similar differential structure, which can produce a longitudinal-bending vibra-
tion when applying a DC field, can reject in-phase vibration noise and enhance the out-
of-phase ME voltage signal simultaneously [54]. We note here that in [54] the authors
found that a ME flux gate sensor excited under a non-resonant high frequency field could
perform better detection ability. As shown in Figure 7f, the relative change of the ME volt-
age output signal in response to a LOD of 1 nT is around 0.2% and the output signal can
return to the reference level during the repeated test cycles when choosing a non-resonant
frequency of 48.5 kHz [64].
Performance summary of some typical magnetoelectric sensors was given in Table 2.
Table 3 further compares the LoD of passive ME sensors with some commercially availa-
ble magnetometers, i.e., magnetoresistive sensors, giant magneto-impedance sensors,
fluxgate sensors, optically pumped magnetometers and SQUID magnetometers. As it can
be seen in Table 3, ME sensor shows comparable and competitive performance with these
products. Specifically, the low power consumption and high detection ability are signifi-
cant advantages for ME sensors, while vibration interference still now greatly limits the
engineering applications. On the other hand, piezoelectric materials are normally suscep-
tible to the working temperature and the temperature stability of ME sensors is also a
critical issue. For example, Burdin et al. compared the temperature dependence of the res-
onant magnetoelectric effect in several kinds of ME composites and showed that the
widely studied PZT-Metglas ME sensor can only work in a narrow temperature range of
0 °C to +50 °C [67].
Actuators 2021, 10, 109 11 of 23
Table 2. Performance summary of typical magnetoelectric sensors.
Composition Working Mode Sensing Mode
Low-frequency mag-
netic field sensing
Metglas/Mn-PMNT [55] Longitudinal vibration (Multi-L-
T) Passive sensing 0.87 pT/
@ 30 Hz
Metglas/PMN-PT [33] Longitudinal vibration (Multi-
push-pull) Passive sensing 5.1 pT/
@ 1 Hz
Metglas/PMN-PZT [61] Longitudinal vibration (L-T) Active Modulation 33 pT/
@ 0.1 Hz
Resonant magnetic
field sensing
Metglas/ LiNbO3 [48] bending mode Direct Sensing 92 fT/√Hz
FeCoSiB/(Pt)/AlN [47] bending mode Direct Sensing 400 fT/√Hz
Metglas/PMN-PZT [8] Longitudinal vibration (L-T) Direct Sensing 123 fT/√Hz
DC magnetic field
sensing
langatate-Metglas [65] bending mode Nonlinear ME effect 10 nT
Metglas/PMN-PZT [9] Longitudinal vibration (L-T) Linear ME effect 1 nT
FeCoSiB/(Pt)/AlN [26] Lateral vibration Delta-E effect 0.8 nT
FeCoSiB/(Pt)/AlN [51] Lateral vibration Delta-E effect 0.4 nT
Table 3. Performance Comparison with commercially available magnetometer for 1 Hz magnetic field sensing.
Magnetometer Working
Temperature
Power
Consumption (mW)
Typical Size
@
(pT/
√
) Limitations
ME sensor [33] 0 °C to +50 °C <1 80 mm × 10 mm
@ ME composites 5.1 Vibration
interference
Magnetoresistive sensor
−40 °C to +125 °C ~0.02 6 mm × 5 mm × 1.5 mm
@ sensing element 100 Low
sensitivity
Giant magneto-
impedance
sensor −20 °C to +60 °C 75 35 mm × 11 mm × 4.6 mm
@ sensing element 15–25 Low
sensitivity
Fluxgate magnetometer
−40 °C to +70 °C 350 ø100 mm × 125 mm
@ system size 2–6 Power
consumption
Optically pumped magne-
tometer −35 °C to +50 °C >12,000 175 cm × 28 cm × 28 cm
@ system size 4 Complex setup
SQUID magnetometer [68]
<−196 °C >1000 12.5 mm× 12.5 mm
@ chip size <0.005 Cooling
Estimated from the data in ref. [65]; Based on commercial product TMR9001 in MultiDimension Technology Co., Ltd.
(Zhangjiagang Free Trade Zone, Jiangsu Province, China); Based on commercial product MI-CB-1DH in AICHI STEEL
CORPORATION (Tōkai city, Aichi Prefecture, Japan); Based on commercial product Mag03 from Bartington Instru-
ments Ltd (Witney, Oxon, OX28 4GG United Kingdom).; Based on commercial product G882 marine magnetometer
from GEOMETRICS, INC (San Jose, CA, USA).
4. Engineering Applications of ME Sensors
As we summarized in Tables 2 and 3, ME sensors show competitive performance
with commercial optically pumped magnetometers, giant magneto-impedance sensors
and fluxgate magnetometers. In this regard, a large number of works that utilize ME sen-
sors for magnetic field sensing have been published and various applications have been
implemented. In this section, we will overview current advances in sensing applications
of ME sensors.
4.1. .Magnetic Target Detection and Localization
A metallic material can be magnetized by the geomagnetic field along each of its three
orthogonal directions, which in turn generates three magnetic signature vectors. In 2012,
Shen et al. proposed a triple-axis magnetic anomaly detection system based on
Metglas/Pb(Zr,Ti)O3/Metglas ME sensors. To compare the performance of ME sensor with
widely used fluxgate sensor, they placed a tri-axial ME sensor, a tri-axial piezoelectric
sensor (PE) and a fluxgate sensor on the ground in a line with the same closest path ap-
proach (CPA) as shown in Figure 8a. The obtained signals are given in Figure 8b. It can be
seen there was little vibration or acoustic contribution to the tri-axial ME sensor at 2.5 s,
and the response amplitude of ME sensor was obviously higher than that of the fluxgate
Actuators 2021, 10, 109 12 of 23
sensor [69]. Then, they also modeled the magnetic field of metallic objects as a magnetic
dipole and designed a ME sensor array to analyze the magnetic signature. The orientation
and velocity of target were presented in the form of a characteristic magnetic waveform,
which provided the basis for magnetic anomaly detection and identification [70].
Figure 8. (a) Photograph of the vehicle detection system setup; (b) sensor output signals in terms of the X (blue curve), Y
(red curve) and Z (green curve) component in the ME sensor (top), PE sensor (middle) and fluxgate sensor (bottom) [69].
With respect to magnetic object localization, Xu et al. constructed a magnetic gradient
meter consisting of eight measuring points using 2-1 type ME sensors to locate the mag-
netic source in a 3-D cube [71]. Their experimentally obtained data that successfully
yielded 3-D vector outputs representing the distribution of the magnetic flux lines of the
tested source placed in the center of the cube. Based on a 1-1 type ME sensor, Chu et al.
also proposed a 2-D magnetic positioning and sketching system consisting of a 1-D ME
sensor array as shown in Figure 9a. Experimental results concerning the positioning of an
iron ball with a diameter of 5 mm are given in Figure 9b. The localization error was
roughly estimated as 2.4 cm in a scanning area of 70 cm (in length) × 50 cm (in height). In
addition, the posture and the length-diameter ratio of a metal bar could be also success-
fully recognized via the system [72].
Figure 9. (a) Prototype of the measurement setup using 1-D ME sensor array and imaging system. (b) the positioning
result of an Fe-ball [72].
Actuators 2021, 10, 109 13 of 23
4.2. Geomagnetic Field Sensing
When it comes to applications, e.g., vehicle detection, maritime navigation and aero-
nautical magnetic exploration, it is necessary to obtain the orientation of a DC (or AC)
magnetic field. In 2007, Zhai et al. measured the value of both the Earth’s magnetic field
and its inclination based on Metglas/piezoelectric-fiber ME laminates [73]. This ME sensor
was working under an active mode without any bias field and it was found that DC mag-
netic field variations of less than 10−9 T and angular inclinations of less than 10−5 deg can
be detected. The voltage output measured when the ME sensor was rotated in the Earth’s
plane is shown in Figure 10. It could be clearly seen that the maximum voltage output
could reach −100 mV when the sensitive direction of the ME sensor was oriented along
the geomagnetic field in the Earth’s plane, but when the sensor was oriented along the
east and west directions in this plane, its output was essentially 0 mV, implying a high
single-axis directivity and the potential application in geomagnetic navigation.
In 2013, Duc et al. presented an integrated two-dimensional geomagnetic device
which consisted of two one-dimensional magnetoelectric sensors arranged in an orthogo-
nal direction. They also finished the circuit design and hardware implementation (Figure
10b). The obtained accuracy of angle measurement reached as low as ±0.5° (Figure 10c)
and the spatial angles can be automatically computed while rotating the sensor module
[74].
Figure 10. (a)Output voltage from the magnetoelectric sensor when it is rotated in the Earth’s plane [73]. (b) Photograph
of the experimental rotation system and (c) the accuracy measurement of spatial angles [74].
In 2018, Pourhosseini et al. presented a multi-terminal hexagonal-framed magnetoe-
lectric composite (HFMEC) to realize the same goal as mentioned above. The hexagonal-
framed ME composite (HFMEC) was fabricated by using three thickness poled [011]-ori-
ented PMN-PT single crystal fibers (Pb(Mg1/3Nb2/3)O3-PbTiO3) sandwiched between two
hexagonal-framed FeBSi alloys (Metglas).
Figure 11. (a) The structure of the HFMEC. (b) ME coupling response of the HFMEC as a function of the DC magnetic-
field, the signal exhibits a “V” shape and the knee point of the curve reveals the geomagnetic field [75].
Actuators 2021, 10, 109 14 of 23
According to the threefold symmetric ME voltage responses of the HFMEC, a valid
formula for calculating the angular direction of the magnetic field was given and a high
angular resolution of about 0.1° was experimentally verified. In addition, the obtained
“V” shape as shown in Figure 11b in the output signal could also be used to determine the
absolute geomagnetic field intensity [75].
4.3. Current Sensing and Non-Destructive Detection
With respect to power grids and chips monitoring and protection, current detection
is urgently needed. Based on Ampere’s law, non-contact measurement is usually achieved
by magnetic sensors, such as current transformers (CTs), Hall devices, SQUIDs, and mag-
netoresistance elements. However, CTs and Hall devices are limited by low sensitivity
and narrow dynamic range; SQUIDs must work at extremely low temperatures and the
inherent 1/f noise also restricts the magnetoresistance elements’ sensitivity [76]. In this
case, researchers tried to achieve high-sensitivity current detection by using ME sensors.
For example, Dong et al. first proposed a ME current sensor based on a ring-type ME
composite consisting of a Pb(Zr,Ti)O3 core laminated between two Terfenol-D layers and
operated in a circumferentially magnetized and circumferentially polarized mode, which
featured a stable frequency response ranging from 0.5 Hz to 2 kHz [77]. However, it was
difficult to apply the magnetic bias along its easily magnetized axes with respect to a ring
type Terfenol-D and it was highly desired to design a self-biased ME current sensor. Then
Zhang et al. further proposed to construct a Fe73.5Cu1Nb3Si13.5B9 air-gapped high-permea-
bility flux concentrator to enhanced the ME response of SmFe2/PZT/SmFe2 self-biased
magnetoelectric (ME) laminate as shown in Figure 12a. In Figure 12b, they compared the
detection ability at 119.75 kHz (resonance) and 1kHz (off-resonance), respectively. A lin-
ear output response was obtained for both cases, and the output sensitivity measured at
resonance was roughly one order of magnitude higher [78].
Oil and natural gas pipelines are normally buried underground or distributed under
the sea, and are thus susceptible to corrosion and pressure damage. The eddy current test-
ing (ECT) technology is widely to monitor the working status of those pipelines [79]. In
2021, Chu et al. constructed an ECT probe, by integrating the exciting coil and a 1-1 type
ME sensor. During the experiment, they set the lift-off distance to 3 mm. Figure 12a shows
the measurement result for an aluminum pipeline with a 10 × 1 × 1 mm3 crack placed in
the center. The scanning signal displayed an obvious response to this crack. In order to
further compare the crack detection and localization performance, authors also separately
tested an aluminum pipeline and a steel pipeline both with a one-dimensionally distrib-
uted cracks labelled 1# and 2#, respectively. Specific response to different cracks was ob-
viously obtained and the location was also identified for both two cases as shown in Fig-
ure 12d. It should be noted here the power consumption of ME ECT probe was only 0.625
μW, which represents a 2–3 orders of magnitude improvement compared with magneto-
resistive (MR) sensors and represented a crucial step towards online low-power monitor-
ing of pipeline cracks [80].
Actuators 2021, 10, 109 15 of 23
Figure 12. (a) Schematic and photograph of the current-sensing device; (b) Comparison of the sen-
sitivities measured at low frequency and resonant frequency [78]. (c) Measured ME voltage re-
sponse with respect to an aluminum pipe with a 10 mm crack in the center (the inset highlighted
by dashed line showed the structure of the ME ECT probe). (d) Crack identifying results for one-
dimensionally distributed cracks labelled 1# and 2# for an aluminum pipeline and steel pipeline,
respectively [80].
4.4. Velocity and Displacement Sensing
It is of great significance to monitor the position and speed of crankshafts in the field
of industrial automation. In 2018, Wu et al. fabricated an angle speed sensor based on a
ME sensor, which consisted of four PZT/FeGa/PZT magnetoelectric (ME) laminate com-
posites, a closed magnetic circuit composed by a magnetic accumulation (MA) ring and
four magnetic accumulation arcs with magnetostrictive layer embedded into the U-
shaped slots, a multi-polar magnetic ring (MPMR), and a holder composed of a shaft and
shells as shown in Figure 13a. The rotating shaft drove the MPMR, which applied alter-
nating magnetic field to ME sensor. As the speed enlarged, the output voltage increased
at first, and then reached to stable values, as shown in Figure 13b. In addition, the meas-
ured frequency of the generated alternating magnetic field presented a linear response as
a function of the rotation speed, as given in Figure 13c. The authors accordingly concluded
the sensor could distinguish a small step-change rotational angle of 0.2° under a rotational
speed of 120 r/min [81].
In 2021, Lu et al. found that during the rotation of the gear, the change of the perme-
ability between the tooth and the air will disturb the magnetic field produced by the mag-
net, which can be used to measure the rotational speed of gear as well. The proposed
system was composed of a ME sensor, a gear and a permanent magnet, as shown in Figure
13d, When the gear was rotating, the convex and concave teeth of the gear alternately
passed, which generated alternating magnetic flux density. As shown in Figure 13e, the
spatial magnetic field varied as the gear rotation angle changed. When the convex tooth
was parallel to the permanent magnet, high permeability of gear led to large magnetic
field at the position of ME sensor. On the contrary, the low permeability of air led to a
small magnetic field. Then, they determined the best distance of three parts of the system
in Figure 13d, and obtained the curve of measured speed (Gs) as a function of setting speed
(Sset), as shown in Figure 13f [82].
Actuators 2021, 10, 109 16 of 23
Micro-displacement sensing is crucial for precise positioning applications, including
micro-manufacturing and biological engineering. Without the obstruction of fringe effects
occurred in capacitive sensors, and the limitations of complicated optical components re-
quired in optical sensors, the ME effect was recently proposed for micro displacement
measurement. In 2021, Yang et al. fabricated a ME sensor based on Terfenol-D/PZT com-
posites [83]. The change of displacement led to the change of air gap (δ(d)), which brought
variation to the magnetic field applied to Terfenol-D by the permanent magnet and thus
enabled an ME voltage output. Then, they investigated the dynamic displacement ampli-
tude and static position measurement performance of the prototype. The ME sensor
achieved its highest sensitivity in the case of resonant excitation and the measured dis-
placement resolution was less than 13.27 nm, which was comparable to and competitive
with commercial laser displacement sensors [83]. A magnetic proximity sensor was also
recently proposed by Pereira et al. based on the ME effect [84].
Figure 13. (a) Schematic view of the packaged sensor; (b) The peak-to-peak values under different rotational speed; (c)
The frequency of the output signal as a function of the rotational speed [81]. (d) Schematic layout of the gear, permanent
magnet and FeCoSiB/Pb(Zr,Ti)O3 sensor; (e) The magnetic flux density with the gear rotates one circle (f) the measurement
speed (Gs) under different rotational speed [82].
4.5. Stress and Strain Measurement
Strain measurement is essential for various applications, such as structural health
monitoring and medical diagnosis. Steel cables are widely used in bridges, ships, mining
and other engineering fields. It is very important to ensure the stress and structure integ-
rity of steel cable. However, the real-time stress monitoring of steel cables is still a difficult
problem. In 2014, Zhang et al. proposed an elasto-magneto-electric (EME) sensor based
on the elasto-magnetic (EM) and magneto-electric (ME) effect, which has advantages of
high sensitivity, fast response, and ease of installation compared with conventional detec-
tion devices [85]. With an exciting coil around the cable, the magnetic field generated by
the current would magnetize the cable and the surrounding area. The stress on the cable
would then change the magnetization strength of the cable, which led to the variation of
the secondary field induced by the cable as shown in Figure 14a. Figure 14b shows their
experimental results, where f was the load value under high stress as measured by the
load cell, and XEME was the normalized result of the signal output of the EME sensor. It
can be seen that a good linear relationship between the stress and the output signal was
obtained [85].
Actuators 2021, 10, 109 17 of 23
In 2020, Chen et al. fabricated a strain sensor based on ME heterostructures, the meas-
urement setup of which was shown in Figure 15a [86]. The frequency shift of ferromag-
netic resonance occurred in NiCo film by applying an elastic strain generated by piezoe-
lectric single crystal PMN-PT was employed to detect the substrate’s strain status. When
excited at 9.4 GHz, the resonant field Hr decreased from 956.65 Oe to 802.83 Oe with the
strain ε increased from 0 με to 700 με, indicating a resonant frequency shift of about 430.7
MHz, as shown in Figure 15b. Accordingly, the calculated strain sensitivity S = (∆f/f)/ε was
determined as 65.46 ppm/με [86]. Considering the resolution of FMR spectra, the limit of
detection of the wireless strain sensor was around 0.54 με, which was comparable with
that of commercial metal-foil sensors that needed connection wires [86].
Figure 14. (a)Structure of the EME sensor; (b) Stress response of EME sensors [85].
Figure 15. (a) the schematic of the EPR cavity and FMR measurement setup; (b) the negative correlation between Hr and
ε of the NiCo film with a static strain range of 0 με–700 με by linear fitting [86].
4.6. Biomagnetic Measurement
Contactless imaging or monitoring of biological entities by using the magnetic field
components of biological currents, has become an emerging field of magnetoelectric sen-
sors [87]. Room temperature working and stringent spatial resolution are typical require-
ments for biomagnetic measurement [88]. Furthermore, the magnetic signals emanating
freely from humans are of very low amplitude compared with the Earth’s magnetic field.
For instance, cardiac magnetic signals are on the order of 10~100 pT, whereas brain mag-
netic signals are typically one to two orders of magnitude lower [89]. Therefore, the mag-
netic detection device needs a super-high sensitivity and wide dynamic range. Figure 16
shows the intensity and frequency distribution of various magnetic signals in human body
and the typical limit of detection of some representative sensors [90]. Compared with
widely used a superconducting quantum interference device (SQUID) requiring liquid
Actuators 2021, 10, 109 18 of 23
nitrogen cooling and optically pumped magnetometers (OPM) suffering from bandwidth
and scalability limitations, ME magnetic field sensors offer passive and thus low-power
detection, high sensitivity, compact structure and also a large dynamic range [91].
In 2016, Hao et al. designed a kind of biomagnetic liver susceptometry (BLS) instru-
ment for assessing the liver iron concentration (LIC) based on ME sensors, which can be
operated under room temperature and in an unshielded environment [88]. In 2020, Lukat
et al. further successfully localized and mapped superparamagnetic iron oxide nanopar-
ticles (SPIONs) with a technique based on ME sensors [92].
Figure 16. Amplitude densities of magnetic signals generated by various sources of the human body [90].
The structure of MSPM is shown in Figure 17a. A rotating plate holder generating
periodic signal was used to enable an easy acquirement of useful signals from the sample
and to remove the environment drift. The permanent magnet and electrically shielded ME
sensor were placed on two sides of the plate holder [92]. Compared with other methods,
this system circumvents the difficulty to separate the magnetic field generated by the sam-
ple under excitation and the magnetic field superposition of the excitation source in mag-
netic particle imaging (MPI) [93] and magnetic particle mapping (MPM) [94]. The mag-
netic signal originating from different concentrations of SPIONs was measured to deter-
mine the performance of the system. As shown in Figure 17b, the signal strength linearly
depended on concentrations of SPIONs and the limit of detectable iron content was about
20 μg. With respect to the spatial resolution, different samples can be well distinguished
in a smallest distance of 5 mm as shown in Figure 17c [92].
Actuators 2021, 10, 109 19 of 23
Figure 17. (a) Setup for measuring distributions of SPIONs. (b) Maximum magnetic field amplitude for different iron
content measurements. (c) Place the sample in different grooves (top), the measured field distributions of the two sample,
and the red line in the field distribution shows the location of the grooves (middle), the particle distribution reconstructed
from these measurements (bottom) [92]. (d) Measurement setup of cardiological magnetic detection. (e) Averaged result
of the R-wave measurement [95].
Reermann et al. proved the statement that thin-film magnetoelectric sensors could be
used for heart magnetic measurements. They successfully measured the R-wave signals
from the heart of a volunteer in a shielded room as shown in Figure 17d, and the inset
highlights the layout of the ME sensor used. The sensor was placed at a distance of about
10 mm above the skin. The result of two times measurement was shown in Figure 17e.
After several-times average of the collected data, a specific peak corresponding to the R-
wave of a human heart was obviously observed [95].
5. Future Outlook
In this review article, we have introduced the advances in ME sensors and their en-
gineering applications. Here, we end with some perspectives that we suggest should be
addressed in the coming future. First, continued efforts should be made to further enhance
the ME coefficient for both thin-film and bulk ME composites by improving the interfacial
stress transfer, looking for better component materials and optimizing the structure pa-
rameters. Considering the specific application for weak magnetic field sensing, the noise
contribution in ME composites should be emphasized as well. Secondly, the realization of
quasi-static and DC magnetic sensing with satisfactory sensitivity based on ME compo-
sites remains a big challenge. Compared with widely studied passive ME sensors, active
ones could be one possible scheme to address this problem. In this case, studies about the
packaging technology to keep a low damping and the corresponding circuit design should
be a priority. In addition, methods to eliminate the high influence from environment vi-
bration should be also considered in a bid to realize engineering applications of ME sen-
sors in relatively complex environment. Finally, we appeal to the industrial community
to get involved in the development of ME sensors for the applications discussed in this
article. Besides the LoD, other critical parameters as mentioned at the beginning of Section
3 should be studied as well. For example, efforts could be made to further improve the
Actuators 2021, 10, 109 20 of 23
temperature stability, to enlarge the dynamic range and to guarantee the final linearity of
a ME sensor. This may create new dimensions for ME community and facilitate the indus-
trialization of ME products.
Author Contributions: Conceptualization, Z.C., J.G.; writing—original draft preparation, J.G., Z.J.;
writing—review and editing, Z.C., Y.S.; visualization, Z.J., Z.C.; supervision, S.Z., Y.S.; funding ac-
quisition, Z.C., Z.M. All authors have read and agreed to the published version of the manuscript.
Funding: This research work was funded by Fundamental scientific research business expenses of
central universities (3072021CFJ0501).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the
design of the study; in the collection, analyses, or interpretation of data; in the writing of the manu-
script, and in the decision to publish the results.
References
1. Ma, J.; Hu, J.; Li, Z.; Nan, C.W. Recent progress in multiferroic magnetoelectric composites: From bulk to thin films. Adv. Mater.
2011, 23, 1062–1087.
2. Hu, J.-M.; Nan, T.; Sun, N.X.; Chen, L.-Q. Multiferroic magnetoelectric nanostructures for novel device applications. MRS Bull.
2015, 40, 728–735.
3. Hu, J.-M.; Duan, C.-G.; Nan, C.-W.; Chen, L.-Q. Understanding and designing magnetoelectric heterostructures guided by
computation: Progresses, remaining questions, and perspectives. NPJ Comput. Mater. 2017, 3, 1–21.
4. Kleemann, W. Multiferroic and magnetoelectric nanocomposites for data processing. J. Phys. D Appl. Phys. 2017, 50, 223001.
5. Lawes, G.; Srinivasan, G. Introduction to magnetoelectric coupling and multiferroic films. J. Phys. D Appl. Phys. 2011, 44, 243001.
6. Palneedi, H.; Annapureddy, V.; Priya, S.; Ryu, J. Status and Perspectives of Multiferroic Magnetoelectric Composite Materials
and Applications. Actuators 2016, 5, 9.
7. Chu, Z.; PourhosseiniAsl, M.; Dong, S. Review of multi-layered magnetoelectric composite materials and devices applications.
J. Phys. D Appl. Phys. 2018, 51, 243001.
8. Chu, Z.; Shi, H.; Shi, W.; Liu, G.; Wu, J.; Yalng, J.; Dong, S. Enhanced Resonance Magnetoelectric Coupling in (1-1) Connectivity
Composites. Adv. Mater. 2017, 29, 1606022.
9. PourhosseiniAsl, M.; Gao, X.; Kamalisiahroudi, S.; Yu, Z.; Chu, Z.; Yang, J.; Lee, H.-Y.; Dong, S. Versatile power and energy
conversion of magnetoelectric composite materials with high efficiency via electromechanical resonance. Nano Energy 2020, 70,
104506.
10. Nan, C.-W.; Bichurin, M.I.; Dong, S.; Viehland, D.; Srinivasan, G. Multiferroic magnetoelectric composites: Historical
perspective, status, and future directions. J. Appl. Phys. 2008, 103, 031101.
11. Fiebig, M. Revival of the magnetoelectric effect. J. Phys. D Appl. Phys. 2005, 38, R123–R152.
12. Zhai, J.; Xing, Z.; Dong, S.; Li, J.; Viehland, D. Magnetoelectric Laminate Composites: An Overview. J. Am. Ceram. Soc. 2008, 91,
351–358.
13. Wang, Y.; Li, J.; Viehland, D. Magnetoelectrics for magnetic sensor applications: Status, challenges and perspectives. Mater.
Today 2014, 17, 269–275.
14. Gutierrez, J.; Lasheras, A.; Martins, P.; Pereira, N.; Barandiaran, J.M.; Lanceros-Mendez, S. Metallic Glass/PVDF Magnetoelectric
Laminates for Resonant Sensors and Actuators: A Review. Sensors 2017, 17, 1251.
15. Sreenivasulu, G.; Qu, P.; Petrov, V.; Qu, H.; Srinivasan, G. Sensitivity Enhancement in Magnetic Sensors Based on Ferroelectric-
Bimorphs and Multiferroic Composites. Sensors 2016, 16, 262
16. Annapureddy, V.; Palneedi, H.; Hwang, G.-T.; Peddigari, M.; Jeong, D.-Y.; Yoon, W.-H.; Kim, K.-H. Magnetic energy harvesting
with magnetoelectrics: An emerging technology for self-powered autonomous systems. Sustain. Energy Fuels 2017, 1, 2039–2052.
17. Vopson, M.M. Fundamentals of Multiferroic Materials and Their Possible Applications. Crit. Rev. Solid State Mater. Sci. 2015, 40,
223–250.
18. Gao, X.; Yan, Y.; Carazo, A.V.; Dong, S.; Priya, S. Low-temperature cofired unipoled multilayer piezoelectric transformers. IEEE
Trans. Ultrason. Ferroelectr. Freq. Control 2017, 65, 513–519, doi:10.1109/tuffc.2017.2785356.
19. Yang, N.; Wu, H.; Wang, S.; Yuan, G.; Zhang, J.; Sokolov, O.; Bichurin, M.I.; Wang, K.; Wang, Y. Ultrasensitive flexible
magnetoelectric sensor. APL Mater. 2021, 9, 021123.
20. Dong, S.; Liu, J.-M.; Cheong, S.-W.; Ren, Z. Multiferroic materials and magnetoelectric physics: Symmetry, entanglement,
excitation, and topology. Adv. Phys. 2015, 64, 519–626.
21. Astrov, D.N. Magnetoelectric Effect In Chromium Oxide. Sov. Phys. Jetp 1961, 13, 729–733.
Actuators 2021, 10, 109 21 of 23
22. Schmid, H. Multi-ferroic magnetoelectrics. Ferroelectrics 1994, 162, 317–338.
23. Catalan, G.; Scott, J.F. Physics and Applications of Bismuth Ferrite. Adv. Mater. 2009, 21, 463–2485.
24. Yamauchi, K.; Picozzi, S. Orbital degrees of freedom as origin of magnetoelectric coupling in magnetite. Phys. Rev. B 2012, 85,
085131.
25. Pantinakis, A.; Jackson, D.A. High-sensitivity low-frequency magnetometer using magnetostrictive primary sensing and
piezoelectric signal recovery. Electron. Lett. 1986, 22, 737–738.
26. 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.
27. Ryu, J.; Priya, S.; Uchino, K.; Kim, H.-E. Magnetoelectric Effect in Composites of Magnetostrictive and Piezoelectric Materials.
J. Electroceramics 2002, 8, 107–119.
28. Dong, S.; Li, J.-F.; Viehland, D. Ultrahigh magnetic field sensitivity in laminates of TERFENOL-D and Pb(Mg1/3Nb2/3)O3–
PbTiO3 crystals. Appl. Phys. Lett. 2003, 83, 2265–2267.
29. Dong, S.; Zhai, J.; Bai, F.; Li, J.-F.; Viehland, D. Push-pull mode magnetostrictive/piezoelectric laminate composite with an
enhanced magnetoelectric voltage coefficient. Appl. Phys. Lett. 2005, 87, 062502.
30. Dong, S.; Zhai, J.; Li, J.; Viehland, D. Near-ideal magnetoelectricity in high-permeability magnetostrictive/piezofiber laminates
with a (2-1) connectivity. Appl. Phys. Lett. 2006, 89, 252904.
31. Gao, J.; Shen, L.; Wang, Y.; Gray, D.; Li, J.; Viehland, D. Enhanced sensitivity to direct current magnetic field changes in
Metglas/Pb(Mg1/3Nb2/3)O3–PbTiO3 laminates. J. Appl. Phys. 2011, 109, 074507.
32. Li, M.; Gao, J.; Wang, Y.; Gray, D.; Li, J.; Viehland, D. Enhancement in magnetic field sensitivity and reduction in equivalent
magnetic noise by magnetoelectric laminate stacks. J. Appl. Phys. 2012, 111, 104504.
33. Wang, Y.; Gray, D.; Betrry, D.; Gao, J.; Li, M.; Li, J.; Viehland, D. An extremely low equivalent magnetic noise magnetoelectric
sensor. Adv. Mater. 2011, 23, 4111–4114.
34. Li, M.; Berry, D.; Das, J.; Gray, D.; Li, J.; Viehland, D. Enhanced Sensitivity and Reduced Noise Floor in Magnetoelectric
Laminate Sensors by an Improved Lamination Process. J. Am. Ceram. Soc. 2011, 94, 3738–3741.
35. Gao, J.; Das, J.; Xing, Z.; Li, J.; Viehland, D. Comparison of noise floor and sensitivity for different magnetoelectric laminates. J.
Appl. Phys. 2010, 108, 084509.
36. Shi, Z.; Nan, C.W.; Zhang, J.; Cai, N.; Li, J.-F. Magnetoelectric effect of Pb(Zr,Ti)O3 rod arrays in a (Tb,Dy)Fe2/epoxy medium.
Appl. Phys. Lett. 2005, 87, 012503.
37. Ma, J.; Shi, Z.; Nan, C.W. Magnetoelectric Properties of Composites of Single Pb(Zr,Ti)O3 Rods and Terfenol-D/Epoxy with a
Single-Period of 1-3-Type Structure. Adv. Mater. 2007, 19, 2571–2573.
38. Palneedi, H.; Choi, S.-Y.; Kim, G.; Annapureddy, V.; Maurya, D.; Priya, S.; Lee, K.J.; Chung, S.; Kang, S.L.; Ryu, J. Tailoring the
Magnetoelectric Properties of Pb(Zr,Ti)O3 Film Deposited on Amorphous Metglas Foil by Laser Annealing. J. Am. Ceram. Soc.
2016, 99, 2680–2687.
39. Palneedi, H.; Maurya, D.; Kim, G.-Y.; Priya, S.; Kang, S.-J.L.; Kim, K.-H.; Choi, S.-Y.; Ryu, J. Enhanced off-resonance
magnetoelectric response in laser annealed PZT thick film grown on magnetostrictive amorphous metal substrate. Appl. Phys.
Lett. 2015, 107, 012904.
40. Palneedi, H.; Yeo, H.G.; Hwang, G.-T.; Annapureddy, V.; Kim, J.-W.; Choi, J.-J.; Trolier-McKinstry, S.; Ryu, J. A flexible, high-
performance magnetoelectric heterostructure of (001) oriented Pb(Zr0.52Ti0.48)O3 film grown on Ni foil. Apl Mater. 2017, 5,
096111.
41. Palneedi, H.; Maurya, D.; Kim, G.Y.; Annapureddy, V.; Noh, M.S.; Kang, C.Y.; Kim, J.W.; Choi, J.J.; Choi, S.Y.;
Chung, S.Y.; et al. Unleashing the Full Potential of Magnetoelectric Coupling in Film Heterostructures. Adv. Mater. 2017, 29,
1605688.
42. Srinivasan, G.; Rasmussen, E.T.; Gallegos, J.; Srinivasan, R.; Bokhan, Y.I.; Laletin, V.M. Magnetoelectric bilayer and multilayer
structures of magnetostrictive and piezoelectric oxides. Phys. Rev. B 2001, 64, 214408.
43. Zhai, J.; Dong, S.; Xing, Z.; Li, J.; Viehland, D. Giant magnetoelectric effect in Metglas/polyvinylidene-fluoride laminates. Appl.
Phys. Lett. 2006, 89, 083507.
44. Jin, J.; Lu, S.-G.; Chanthad, C.; Zhang, Q.; Haque, M.A.; Wang, Q. Multiferroic Polymer Composites with Greatly Enhanced
Magnetoelectric Effect under a Low Magnetic Bias. Adv. Mater. 2011, 23, 3853.
45. Sreenivasulu, G.; Fetisov, L.Y.; Fetisov, Y.K.; Srinivasan, G. Piezoelectric single crystal langatate and ferromagnetic composites:
Studies on low-frequency and resonance magnetoelectric effects. Appl. Phys. Lett. 2012, 100, 052901.
46. Kirchhof, C.; Krantz, M.; Teliban, I.; Jahns, R.; Marauska, S.; Wagner, B.; Knöchel, R.; Gerken, M.; Meyners, D.; Quandt, E. Giant
magnetoelectric effect in vacuum. Appl. Phys. Lett. 2013, 102, 232905.
47. Yarar, E.; Salzer, S.; Hrkac, V.; Piorra, A.; Höft, M.; Knöchel, R.; Kienle, L.; Quandt, E. Inverse bilayer magnetoelectric thin film
sensor. Appl. Phys. Lett. 2016, 109, 022901.
48. Turutin, A.V.; Vidal, J.V.; Kubasov, I.V.; Kislyuk, A.M.; Malinkovich, M.D.; Parkhomenko, Y.N.; Kobeleva, S.P.; Pakhomov,
O.V.; Kholkin, A.L.; Sobolev, N.A. Magnetoelectric metglas/bidomain y + 140°-cut lithium niobate composite for sensing fT
magnetic fields. Appl. Phys. Lett. 2018, 112, 262906.
49. Greve, H.; Woltermann, E.; Quenzer, H.-J.; Wagner, B.; Quandt, E. Giant magnetoelectric coefficients in (Fe90Co10)78Si12B10-
AlN thin film composites. Appl. Phys. Lett. 2010, 96, 182501.
Actuators 2021, 10, 109 22 of 23
50. Jovičević Klug, M.; Thormählen, L.; Röbisch, V.; Toxværd, S.D.; Höft, M.; Knöchel, R.; Quandt, E.; Meyners, D.; McCord, J.
Antiparallel exchange biased multilayers for low magnetic noise magnetic field sensors. Appl. Phys. Lett. 2019, 114, 192410.
51. Nan, T.; Hui, Y.; Rinaldi, M.; Sun, N.X. Self-biased 215 MHz magnetoelectric NEMS resonator for ultra-sensitive DC magnetic
field detection. Sci. Rep. 2013, 3, 1985.
52. Tu, C.; Chu, Z.-Q.; Spetzler, B.; Hayes, P.; Dong, C.-Z.; Liang, X.-F.; Chen, H.-H.; He, Y.-F.; Wei, Y.-Y.; Lisenkov, I.; et al.
Mechanical-Resonance-Enhanced Thin-Film Magnetoelectric Heterostructures for Magnetometers, Mechanical Antennas,
Tunable RF Inductors, and Filters. Materials 2019, 12, 2259.
53. Nan, T.; Lin, H.; Gao, Y.; Matyushov, A.; Yu, G.; Chen, H.; Sun, N.; Wei, S.; Wang, Z.; Li, M.; et al. Acoustically actuated ultra-
compact NEMS magnetoelectric antennas. Nat. Commun 2017, 8, 296.
54. Chu, Z.; Shi, H.; Gao, X.; Wu, J.; Dong, S. Magnetoelectric coupling of a magnetoelectric flux gate sensor in vibration noise
circumstance. AIP Adv. 2018, 8, 015203.
55. Fang, C.; Jiao, J.; Ma, J.; Lin, D.; Xu, H.; Zhao, X.; Luo, H. Significant reduction of equivalent magnetic noise by in-plane series
connection in magnetoelectric Metglas/Mn-doped Pb(Mg1/3Nb2/3)O3-PbTiO3laminate composites. J. Phys. D Appl. Phys. 2015,
48, 465002.
56. Liu, Y.; Jiao, J.; Ma, J.; Ren, B.; Li, L.; Zhao, X.; Luo, H.; Shi, L. Frequency conversion in magnetoelectric composites for quasi-
static magnetic field detection. Appl. Phys. Lett. 2013, 103, 212902.
57. Petrie, J.R.; Fine, J.; Mandal, S.; Sreenivasulu, G.; Srinivasan, G.; Edelstein, A.S. Enhanced sensitivity of magnetoelectric sensors
by tuning the resonant frequency. Appl. Phys. Lett. 2011, 99, 043504.
58. Petrie, J.; Mandal, S.; Gollapudi, S.; Viehland, D.; Gray, D.; Srinivasan, G.; Edelstein, A.S. Enhancing the sensitivity of
magnetoelectric sensors by increasing the operating frequency. J. Appl. Phys. 2011, 110, 124506.
59. Ou-Yang, J.; Liu, X.; Zhou, H.; Zou, Z.; Yang, Y.; Li, J.; Zhang, Y.; Zhu, B.; Chetn, S.; Yang, X. Magnetoelectric laminate
composites: An overview of methods for improving the DC and low-frequency response. J. Phys. D Appl. Phys. 2018, 51, 324005.
60. Chu, Z.; Dong, C.; Tu, C.; Liang, X.; Chen, H.; Sun, C.; Yu, Z.; Dong, S.; Sun, N.-X. A low-power and high-sensitivity magnetic
field sensor based on converse magnetoelectric effect. Appl. Phys. Lett. 2019, 115, 162901.
61. Chu, Z.; Yu, Z.; PourhosseiniAsl, M.; Tu, C.; Dong, S. Enhanced low-frequency magnetic field sensitivity in magnetoelectric
composite with amplitude modulation method. Appl. Phys. Lett. 2019, 114, 132901.
62. PourhosseiniAsl, M et al. Enhanced self-bias magnetoelectric effect in locally heat-treated ME laminated composite. Appl. Phys.
Lett. 2019, 115, 112901.
63. Li, J.; Ma, G.; Zhang, S.; Wang, C.; Jin, Z.; Zong, W.; Zhao, G.; Wang, X.; Xu, J.; Cao, D.; et al. AC/DC dual-mode magnetoelectric
sensor with high magnetic field resolution and broad operating bandwidth. AIP Adv. 2021, 11, 045015.
64. Chu, Z.; Shi, H.; PourhosseiniAsl, M.J.; Wu, J.; Shi, W.; Galo, X.; Yuan, X.; Dong, S. A magnetoelectric flux gate: New approach
for weak DC magnetic field detection. Sci. Rep. 2017, 7, 8592.
65. Burdin, D.A.; Chashin, D.V.; Ekonomov, N.A.; Fetisov, Y.K.; Stashkevich, A.A. High-sensitivity dc field magnetometer using
nonlinear resonance magnetoelectric effect. J. Magn. Magn. Mater. 2016, 405, 244–248.
66. Burdin, D.; Chashin, D.; Ekonomov, N.; Fetisov, L.; Fetisov, Y.; Shamonin, M. DC magnetic field sensing based on the nonlinear
magnetoelectric effect in magnetic heterostructures. J. Phys. D Appl. Phys. 2016, 49, 375002.
67. Burdin, D.A.; Ekonomov, N.A.; Chashin, D.V.; Fetisov, L.Y.; Fetisov, Y.K.; Shamonin, M. Temperature Dependence of the
Resonant Magnetoelectric Effect in Layered Heterostructures. Materials 2017, 10, 1183.
68. Schmelz, M.; Stolz, R.; Zakosarenko, V.; Schönau, T.; Anders, S.; Fritzsch, L.; Mück, M.; Meyer, M.; Meyer, H.-G. Sub-fT/Hz1/2
resolution and field-stable SQUID magnetometer based on low parasitic capacitance sub-micrometer cross-type Josephson
tunnel junctions. Phys. C Supercond. Its Appl. 2012, 482, 27–32.
69. Shen, Y.; Gao, J.; Hasanyan, D.; Wang, Y.; Li, M.; Li, J.; Viehland, D. Investigation of vehicle induced magnetic anomaly by
triple-axis magnetoelectric sensors. Smart Mater. Struct. 2012, 21, 115007.
70. Shen, Y.; Hasanyan, D.; Gao, J.; Wang, Y.; Li, J.; Viehland, D. A magnetic signature study using magnetoelectric laminate sensors.
Smart Mater. Struct. 2013, 22, 095007.
71. Xu, J.; Leung, C.M.; Zhuang, X.; Li, J.; Viehland, D. Spatial magnetic source detection based on active mode magnetoelectric
gradiometer with 2D and 3D configurations. J. Phys. D Appl. Phys. 2020, 53, 365002.
72. Chu, Z.; Shi, W.; Shi, H.; Chen, Q.; Wang, L.; PourhosseiniAsl, M.J.; Xiao, C.; Xie, T.; Dong, S. A 1D Magnetoelectric Sensor
Array for Magnetic Sketching. Adv. Mater. Technol. 2019, 4, 1800484.
73. Zhai, J.; Dong, S.; Xing, Z.; Li, J.; Viehland, D. Geomagnetic sensor based on giant magnetoelectric effect. Appl. Phys. Lett. 2007,
91, 123513.
74. Duc, N.H.; Tu, B.D.; Ngoc, N.T.; Lap, V.D.; Giang, D.T.H. Metglas/PZT-Magnetoelectric 2-D Geomagnetic Device for
Computing Precise Angular Position. IEEE Trans. Magn. 2013, 49, 4839–4842.
75. PourhosseiniAsl, M.J.; Chu, Z.; Gao, X.; Dong, S. A hexagonal-framed magnetoelectric composite for magnetic vector
measurement. Appl. Phys. Lett. 2018, 113, 092902.
76. Zhang, J.; Li, P.; Wen, Y.; He, W.; Yang, J.; Yang, A.; Lu, C.; Li, W. Enhanced sensitivity in magnetoelectric current-sensing
devices with frequency up-conversion mechanism by modulating the magnetostrictive strain. J. Appl. Phys. 2014, 115, 17E505.
77. Dong, S.; Li, J.-F.; Viehland, D. Vortex magnetic field sensor based on ring-type magnetoelectric laminate. Appl. Phys. Lett. 2004,
85, 2307–2309.
Actuators 2021, 10, 109 23 of 23
78. Zhang, J.; Li, P.; Wen, Y.; He, W.; Yang, A.; Lu, C. Packaged current-sensing device with self-biased magnetoelectric laminate
for low-frequency weak-current detection. Smart Mater. Struct. 2014, 23, 095028.
79. Zhao, S.; Sun, L.; Gao, J.; Wang, J.; Shen, Y. Uniaxial ACFM detection system for metal crack size estimation using magnetic
signature waveform analysis. Measurement 2020, 164, 108090.
80. Chu, Z.; Jiang, Z.; Mao, Z.; Shen, Y.; Gao, J.; Dong, S. Low-power eddy current detection with 1-1 type magnetoelectric sensor
for pipeline cracks monitoring. Sens. Actuators A Phys. 2021, 318, 112496.
81. Wu, Z.; Bian, L.; Chen, S. Packaged angle-sensing device with magnetoelectric laminate composite and magnetic circuit. Sens.
Actuators A Phys. 2018, 273, 232–239.
82. Lu, C.; Zhu, R.; Yu, F.; Jiang, X.; Liu, Z.; Dong, L.; Hua, Q.; Ou, Z. Gear rotational speed sensor based on FeCoSiB/Pb(Zr,Ti)O3
magnetoelectric composite. Measurement 2021, 168, 108409.
83. Yang, Y.; Yang, B. Displacement Sensor with Nanometric resolution Based on Magnetoelectric Effect. IEEE Sens. J. 2021, 21,
12084–12091.
84. Pereira, N.; Lima, A.C.; Correia, V.; Perinka, N.; Lanceros-Mendez, S.; Martins, P. Magnetic Proximity Sensor Based on
Magnetoelectric Composites and Printed Coils. Materials 2020, 13, 1729.
85. Zhang, R.; Duan, Y.; Or, S.W.; Zhao, Y. Smart elasto-magneto-electric (EME) sensors for stress monitoring of steel cables: Design
theory and experimental validation. Sensors 2014, 14, 13644–13660.
86. Chen, Y.; Hu, C.; Wang, Z.; Li, Y.; Zhu, S.; Su, W.; Hu, Z.; Zhou, Z.; Liu, M. Wireless strain sensor based on the magnetic strain
anisotropy dependent ferromagnetic resonance. AIP Adv. 2020, 10, 150310
87. Hayes, P.; Klug, M.J.; Toxværd, S.; Durdaut, P.; Schell, V.; Teplyuk, A.; Burdin, D.; Winkler, A.; Weser, R.; Fetisov, Y.; et al.
Converse Magnetoelectric Composite Resonator for Sensing Small Magnetic Fields. Sci. Rep. 2019, 9, 16355.
88. Xi, H.; Qian, X.; Lu, M.-C.; Mei, L.; Rupprecht, S.; Yang, Q.X.; Zhang, Q.M. A Room Temperature Ultrasensitive Magnetoelectric
Susceptometer for Quantitative Tissue Iron Detection. Sci. Rep. 2016, 6, 29740.
89. Wikswo, J.P. SQUID magnetometers for biomagnetism and nondestructive testing: Important questions and initial answers.
IEEE Trans. Appl. Supercond. 1995, 5, 74–120.
90. Zuo, S.; Schmalz, J.; Ozden, M.-O.; Gerken, M.; Su, J.; Niekiel, F.; Lofink, F.; Nazarpour, K.; Heidari, H. Ultrasensitive
Magnetoelectric Sensing System for Pico-Tesla MagnetoMyoGraphy. IEEE Trans. Biomed. Circuits Syst 2020, 14, 971–984.
91. Ripka, P.; Janosek, M. Advances in Magnetic Field Sensors. IEEE Sens. J. 2010, 10, 1108–1116.
92. Lukat, N.; Friedrich, R.-M.; Spetzler, B.; Kirchhof, C.; Arndt, C.; Thormählen, L.; Faupel, F.; Selhuber-Unkel, C. Mapping of
magnetic nanoparticles and cells using thin film magnetoelectric sensors based on the delta-E effect. Sens. Actuators A Phys. 2020,
309, 112023.
93. Gleich, B.; Weizenecker, J. Tomographic imaging using the nonlinear response of magnetic particles. Nature 2005, 435, 1214–
1217.
94. 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, 2086.
95. Reermann, J.; Durdaut, P.; Salzer, S.; Demming, T.; Piorra, A.; Quandt, E.; Frey, N.; Höft, M.; Schmidt, G. Evaluation of
magnetoelectric sensor systems for cardiological applications. Measurement 2018, 116, 230–238.
... The field conversion efficiency can be controlled by applying an additional DC bias magnetic field H to the structure. Such structures are promising for potential applications as magnetic field sensors, signal processing devices, magnetic memory elements, gyrators, controlled inductances and transformers [2][3][4][5]. ...
... In particular, the effects of voltage harmonics and subharmonics generation, alternating magnetic fields frequency mixing, etc. were discovered and studied [6]. Highly sensitive magnetic field sensors [2], a magnetic field spectrum analyzer [7], and frequency doublers [8][9][10] were developed on the basis of these effects. ...
Article
Full-text available
A resonant voltage frequency doubler using the nonlinear magnetoelectric effect in planar heterostructure containing a piezoelectric langatate single crystal placed between two layers of amorphous ferromagnetic Metglas alloy was fabricated and studied. The frequency doubling occurred due to the nonlinear dependence of the Metglas magnetostriction on the magnetic field. The frequency of the voltage generated by the structure was twice higher than the frequency of the excitation field and was equal to the resonant frequency of the structure f0 = 84.6 kHz, which corresponds to the nonlinear ME effect of frequency doubling. A nonlinear magnetoelectric coefficient of 7.1 V/(cm*Oe^2) was obtained at acoustic resonance frequency of the structure due to high acoustic quality factor of ~2820. The power conversion coefficient of the doubler reached ~1% at optimal bias magnetic field of 2 Oe and matched load resistance of 0.8 kΩ.
... The resistance of Co-Ti substituted hexaferrites is reported to be very large that these ferrites can withstand high electric field without producing large leakage current [18,19]. This has opened up a whole new dimension of probable applications of BaM as a single phase multiferroic material in multifunctional device applications like multiple-state memory elements, functional sensors, transducers, and novel memory media [20][21][22]. Despite these advantages, reduction in saturation magnetization has also been observed in previously reported studies [23,24]. ...
... The fitting of the M − H curve at the higher magnetic field end is shown in Fig. 12 and the M s and B values are obtained from it using equation (20). H a and K 1 are then calculated from equation (21). The parameter values thus obtained are tabulated in Table 4. ...
... It is interesting to note that the loss tangent, which is associated with the energy loss during the alignment of electric dipoles during the poling process, it is expected that the sensor will exhibit a lower value of this quantity. This factor may directly interfere with the sensor's sensitivity, as the loss tangent is also linked to the generation of noise in the internal structure of the sensor, impacting its sensitivity [38]. Finally, despite the lack of some information, we can propose that the dielectric measurements of our 1-3 ME sensors reveal typical frequency-dependent behavior with higher dielectric permittivity and loss at low frequencies due to space charge and interfacial polarization, decreasing at higher frequencies due to dipole relaxation processes. ...
Article
Full-text available
This paper presents a novel approach for the fabrication of magnetoelectric composites aimed at enhancing cross-coupling between electrical and magnetic phases for potential applications in intelligent sensors and electronic components. Unlike previous methodologies known for their complexity and expense, our method offers a simple and cost-effective assembly process conducted at room temperature, preserving the original properties of the components and avoiding undesired phases. The composites, composed of PZT fibers, cobalt (CoFe2O4), and a polymeric resin, demonstrate the uniform distribution of PZT-5A fibers within the cobalt matrix, as demonstrated by scanning electron microscopy. Detailed morphological analyses reveal the interface characteristics crucial for determining overall performance. Dielectric measurements indicate stable behaviors, particularly when PZT-5A fibers are properly poled, showcasing potential applications in sensors or medical devices. Furthermore, H-dependence studies illustrate strong magnetoelectric interactions, suggesting promising avenues for enhancing coupling efficiency. Overall, this study lays the basic work for future optimization of composite composition and exploration of its long-term stability, offering valuable insights into the potential applications of magnetoelectric composites in various technological domains.
... Magnetoelectric effects (ME) in composite heterostructures containing ferromagnetic (FM) and piezoelectric (PE) layers are widely used to create high-sensitivity magnetic field and current sensors, controlled electronics devices, low-frequency antennas and harvesters [1][2][3]. These effects arise due to the combination of magnetostriction of the FM layer and piezoelectricity in the PE layer due to mechanical coupling of the layers, and manifest themselves in the form of mutual conversion of magnetic and electric fields. ...
Article
Full-text available
The magnetoelectric (ME) effect was observed and investigated in a planar ring-type heterostructure containing mechanically coupled rings of magnetostrictive amorphous alloy Metglas and a piezoelectric single crystal langatate. The structure was excited by a circumferential ac magnetic field in the presence of a circumferential dc bias magnetic field. Due to the absence of demagnetization and the high acoustic quality factor of the langatate, a high ME field conversion coefficient of α E ≈ 53 V (Oe cm)⁻¹ was obtained at the frequency 221 kHz of the fundamental radial acoustic resonance mode. The structure can be used as a non-contact ac or dc current sensor. The sensitivity of the structure to the ac current was 1.6 V A⁻¹, whereas for the dc current it reached 4.2 V A⁻¹.
... These transducers utilize the magnetoelectric (ME) effect to convert motion velocity into an induced potential output in the coil. These sensors operate without the need for an external power source, instead directly harnessing the mechanical energy of the object being measured and converting it into an electrical signal output [7,8]. Because ME enables the high-sensitivity transformation from a magnetic signal into an electric signal, researchers have found it an attractive option for developing magnetic field sensors and current sensors [9]. ...
Article
Full-text available
The evaluation of blasting vibrations primarily hinges on two physical quantities: velocity and acceleration. A significant challenge arises when attempting to reference the two types of vibration data in relation to one another, such as different types of seismometers, noise, etc., necessitating a method for their equivalent transformation. To address this, a transformation method is discussed in detail with a case study, and equations for the ratio (Ra) of the particle peak velocity (PPV) to the particle peak acceleration (PPA) are proposed. The findings are twofold: (1) The conventional data conversion processes often suffer from low accuracy due to the presence of trend terms and noise in the signal. To mitigate this, the built-in MATLAB function is used for trend term elimination, complemented by a combined approach that integrates CEEMDAN with WD/WDP for noise reduction. These significantly enhance the accuracy of the transformation. (2) This analysis reveals a positive power function correlation between Ra and the propagation distance of the blast vibrations, contrasted by a negative correlation with the maximum charge per delay. Intriguingly, the Ra values observed in pre-splitting blasting operations are consistently lower than those in bench blasting. The established Ra equations offer a rapid, direct method for assessing the transformation between the PPV and PPA, providing valuable insights for the optimization of blasting design.
Article
Magnetoelectric (ME) sensors are widely studied and well suited for current condition monitoring in smart grids due to their high sensitivity, low power consumption, and compact size. However, designing ME sensors that simultaneously achieve low magnetic noise and a large linear range remains challenging. In this work, we propose and systematically study a multilayered magnetoelectric sensor (MLMS). We experimentally demonstrate that the voltage noise is effectively suppressed by connecting the piezoelectric elements of the MLMS in series. Additionally, the magnetic flux concentration effect is weakened by magnetic shielding from the outer Metglas laminates, which increases the optimized bias field to 52 Oe for the MLMS. Consequently, an equivalent magnetic noise as low as 16.7 pT/rtHz at 1 Hz is obtained and an enlarged linear range from 20 pT to 2 mT is achieved. Furthermore, we demonstrate that the proposed MLMS could linearly detect a wide range of power currents from 0.1 mA to 400 A, with nonlinear error and current resolution as low as 0.05% and 0.1 mA, respectively.
Article
Full-text available
In this study, a dual-mode Metglas/Pb(Zr,Ti)O3 magnetoelectric (ME) sensor was prepared for measuring weak magnetic fields. It is interesting to note that this ME sensor can work at alternating current (AC) and direct current (DC) dual-modes with high field resolution. In AC mode, a very accurate AC magnetic field resolution of 0.8 nT was achieved at a mechanical resonance frequency of 72.2 kHz; moreover, the operating frequency band for resolution better than 1 nT is as wide as 3.4 kHz. We proposed a DC bias field perturbation (DBFP) method to detect the DC magnetic field using lock-in amplifier technology. As a result, an ultra-accurate DC field resolution of 0.9 nT with noise power spectral density as low as 100 pT/Hz was obtained in the studied ME sensor via the DBFP method. The dual-mode ME sensor enables simultaneous measurement for DC and AC magnetic fields with wideband and accurate field resolution, which greatly enhances the measurement flexibility and application scope.
Article
Full-text available
Micro-displacement measurement is essential in precision fields, such as precision positioning, micro-vibration control and biological engineering. Conventional sensors have disadvantages of susceptibility to environmental, high cost and difficulty in integration. To address these deficiencies, this paper developed a novel displacement sensor with nanometric resolution based on magnetoelectric effect. Combining equivalent magnetic circuit method and equivalent circuit method for magnetoelectric effect using the nonlinear constitutive parameters, an equivalent circuit model of the proposed sensor is established to analyse and predict the performance of displacement sensor. Then a prototype of sensor based on Terfenol-D/PZT composites is fabricated, and the performance of prototype for dynamic displacement amplitude and static position measurement are tested. The results not only validate the developed equivalent circuit model, but also evidence the potential of magnetoelectric displacement sensor. Especially for dynamic displacement amplitude measurement, the resolution of it is better than 13.27 nm, which is close to or even beyond commercial laser displacement sensor.
Article
Full-text available
Ever-evolving advances in flexible magnetic sensors are promising to fuel technological developments in the fields of touchless human–machine interaction, implantable medical diagnosis, and magnetoreception for artificial intelligence. However, the realization of highly flexible and extremely sensitive magnetic sensors remains a challenge. Here, we report a cost-effective, flexible, and ultra-sensitive heterostructural magnetoelectric (ME) sensor consisting of piezoelectric Pb(Zr0.52Ti0.48)O3 (PZT) thick films and Metglas foils. The flexible sensor exhibits a strong ME coefficient of 19.3 V cm⁻¹ Oe⁻¹ at low frequencies and 280.5 V cm⁻¹ Oe⁻¹ at resonance due to the exceptionally high piezoelectric coefficient d33 ∼ 72 pC N⁻¹ of the constituent PZT thick films. The flexible ME sensor possesses not only ultrahigh sensitivities of 200 nT at low frequencies and 200 pT at resonance but also shows an excellent mechanical endurance. Through 5000 bending cycles (radii of ∼1 cm), the sensors showed no fatigue-induced performance degradation. This ultrasensitive flexible sensor provides a platform capable of sensing and responding to external magnetic fields and will find applications in soft robotics, wearable healthcare monitoring, and consumer electronics.
Article
Full-text available
Wireless strain sensors have received extensive attention owing to their wide application prospects in structural health monitoring, industrial automation, human activity monitoring, and intelligent robotic systems. Here, a wireless strain sensor prototype based on the magnetoelectric heterostructure of ferromagnetic thin films on a piezoelectric substrate has been developed. The ferromagnetic resonance (FMR) frequency of the sensor is strongly dependent on external strain due to the large magnetostriction of the film. The piezoelectric substrate with a programmable voltage has been used as a strain source for the characterization of the wireless strain sensor. The limit of detection of the wireless strain sensor is 0.54 με, which is comparable with that of commercial metal-foil sensors that need connection wires. More importantly, the FMR strain sensor shows a sensitivity of 65.46 ppm/με, indicating more than a 60 fold improvement than that of traditional wireless strain sensors based on patch antenna and RLC resonators whose frequency shift is mainly due to the strain induced dimension change.
Article
Full-text available
MagnetoMyoGraphy (MMG) with superconducting quantum interference devices (SQUIDs) enabled the measurement of very weak magnetic fields (femto to pico Tesla) generated from the human skeletal muscles during contraction. However, SQUIDs are bulky, costly and require working in a temperature-controlled environment, limiting wide-spread clinical use. We introduce a low-profile magnetoelectric (ME) sensor with analog frontend circuitry that has sensitivity to measure pico-Tesla MMG signals at room temperature. It comprises magnetostrictive and piezoelectric materials, FeCoSiB/AlN. Accurate device modelling and simulation are presented to predict device fabrication process comprehensively using the finite element method (FEM) in COMSOL Multiphysics®. The fabricated ME chip with its readout circuit was characterized under a dynamic geomagnetic field cancellation technique. The ME sensor experiment validate a very linear response with high sensitivities of up to 378 V/T driven at a resonance frequency of ${f_{res}}$ = 7.76 kHz. Measurements show the sensor limit of detections of down to 175 pT/Hz at resonance, which is in the range of MMG signals. Such a small-scale sensor has the potential to monitor chronic movement disorders and improve the end-user acceptance of human-machine interfaces.
Article
Full-text available
Magnetic sensors are mandatory in a broad range of applications nowadays, being the increasing interest on such sensors mainly driven by the growing demand of materials required by Industry 4.0 and the Internet of Things concept. Optimized power consumption, reliability, flexibility, versatility, lightweight and low-temperature fabrication are some of the technological requirements in which the scientific community is focusing efforts. Aiming to positively respond to those challenges, this work reports magnetic proximity sensors based on magnetoelectric (ME) polyvinylidene fluoride (PVDF)/Metglas composites and an excitation-printed coil. The proposed magnetic proximity sensor shows a maximum resonant ME coefficient (α) of 50.2 Vcm−1Oe−1, an AC linear response (R2 = 0.997) and a maximum voltage output of 362 mV, which suggests suitability for proximity-sensing applications in the areas of aerospace, automotive, positioning, machine safety, recreation and advertising panels, among others.
Article
Pipeline systems, such as oil and natural gas pipelines, which are normally buried underground or distributed under the sea, are susceptible to corrosion and pressure destruction, and hence timely and reliable nondestructive testing (NDT) is necessary. Conventional nondestructive testing methods are generally based on ultrasonic, eddy current and magnetic flux leakage mechanism. With respect to eddy current testing (ECT) technologies, the current challenge is to develop magnetic sensors with high detection ability, extremely low power consumption and compatible with Internet of Things (IoT). Giant resonant magnetoelectric (ME) coupling has been previously reported in our proposed 1-1 ME composites. Here, we further presented an investigation on the eddy current testing (ECT) using 1-1 type ME sensor. The proposed ECT probe was built by integrating a small exciting coil surrounding the ME sensor. Simulation results indicate ECT based on ME sensors is especially applicable for non-ferromagnetic and low-conductivity materials, e.g. steel. One-dimensionally distributed cracks with varying crack sizes in a steel pipeline was then experimentally identified and located using ECT method. More importantly, the power consumption of our proposed ECT probe is as low as 0.625 uW, which shows 2–3 orders of magnitude improvement compared with other probes based on e.g. magnetoresistive (MR) sensors. Our research here provides a fundamental insight into NDT using ME sensors as a ECT probe and represents a crucial step towards online low-power monitoring of pipeline cracks.
Article
This paper designs a gear rotational speed sensor consisting of a permanent magnet and a FeCoSiB/Pb(Zr,Ti)O3 magnetoelectric composite as a magnetic sensor. The influences of distances between the detection gear, FeCoSiB/Pb(Zr,Ti)O3 composite, and the permanent magnet on output voltage are investigated in detail. In order to achieve a larger output voltage, the FeCoSiB/Pb(Zr,Ti)O3 composite should be closer to the gear. The results show that the proposed sensor has good near-linear to detect rotational speed. When the frequency is the output parameter, the sensitivity of the proposed sensor to speed is determined to be 0.9951 in the range of 10–600 rpm. Furthermore, based on the magnitude of output voltage, the voltage detection sensitivity of the proposed sensor to gear speed is determined to be ~ 1.15 mV/rpm in the range of 10–150 rpm. The research indicates that the magnetoelectric composite has physical feasibility and great potential application prospect in the field of gear speed measurement.
Article
A single-axis alternating current field measurement (ACFM) detection system is proposed for crack size estimation. A single tunneling magneto-resistive (TMR) sensor is used to detect Bz signal, which is able to determine both the length and depth of a crack simultaneously. First, a theoretical analysis is presented to evaluate the crack length and especially depth using Bz signature waveform. The underlying physics principle is supported using a finite element analysis (FEA) method. In the simulation, the cracks with various lengths and depths are analyzed for a given crack width, and corresponding Bzmax values are obtained. Next, a Bzmax characteristic polynomial surface is developed, which can be represented by a fitted polynomial interpolation equation. The crack depth can be inversed by this equation with reference to the measured Bzmax value and crack length. Finally, real ACFM experiments are conducted to demonstrate that the crack lengths and depths can be readily estimated.
Article
Superparamagnetic iron oxide nanoparticles (SPIONs) are an important tool for labeling cells and tissues in many therapeutic and diagnostic applications, such as magnetic resonance imaging (MRI) and magnetic particle imaging (MPI). However, these methods require large and expensive instrumentation. Here we show that our magnetic susceptibility particle mapping (MSPM) system can achieve the detection of magnetic nanoparticles in an inexpensive and small device. The system is based on magnetoelectric (ME) sensors utilizing the ΔE effect in combination with a permanent magnet that is generating a bias field for the sensor and at the same time is magnetizing the SPIONs in the sample. The permanent magnet is placed above the sensor, and the sample is rotated through the gap in between. The magnetized SPIONs in the sample generate an additional magnetic field that can be detected by the ME sensor. The clear novelty of our approach is the use of a rotating sample, generating a periodic signal, which enables an easy separation of the desired signal from the background signal and the possibility to compensate drift, which is commonly observed in ME sensor measurements. With this improvement and the use of a ME sensor that is sensitive for low frequencies the setup is able to measure significantly smaller amounts of magnetic nanoparticles than previous approaches described in the literature and are even able to reconstruct 2D nanoparticle distributions. The noise floor, also referred to as limit of detection (LOD), of this measurement system is around 500 pT/(Hz)1/2. The detection threshold of our MSPM system is 20 μg SPIONs in a volume of 200 mm³ and the spatial resolution is in the range of a few mm. The spatial resolution is determined by reconstructing the particle distribution in the sample layer by solving the inverse problem. To demonstrate the feasibility of the method for detecting living cells, we measured the field distribution originating from SPION-labeled fibroblast cells in an alginate-gelatin matrix, thus demonstrating the potential of our method for biomaterial applications.