ArticlePDF Available

Integrated Magnetics and Magnetoelectrics for Sensing, Power, RF, and Microwave Electronics

IEEE
IEEE Journal of Microwaves
Authors:

Abstract and Figures

As the rapid development of integrated magnetic and magnetoelectric, numerous novel devices including high performance on-chip transformers, inductors, filters, antennas, and sensors with unique advantages in power efficiency, size and tunability, etc. have been demonstrated. In this review, an overview of the development of magnetism and magnetoelectric will be firstly given. The conceptual illustration and materials used in integrated magnetoelectric will then be presented. Selections of on-chip devices from literatures will be shown to exemplify the integrated magnetic and magnetoelectric applications. Finally, the prospect and the direction of the future research will be discussed in the conclusion.
Content may be subject to copyright.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
Received 1 May 2021; revised 1 August 2021; accepted 23 August 2021.
Digital Object Identifier 10.1109/JMW.2021.3109277
Integrated Magnetics and Magnetoelectrics for
Sensing, Power, RF, and Microwave Electronics
YIFAN HE ,BINLUO , AND NIAN-XIANG SUN (Fellow, IEEE)
(Invited Paper)
Department of Electrical and Computer Engineering, Northeastern University, Boston, MA 02115 USA
CORRESPONDING AUTHOR: Nian-Xiang Sun (e-mail: n.sun@northeastern.edu).
(Yifan He and Bin Luo contributed equally to this work.)
ABSTRACT As the rapid development of integrated magnetic and magnetoelectric, numerous novel devices
including high performance on-chip transformers, inductors, filters, antennas, and sensors with unique
advantages in power efficiency, size and tunability, etc. have been demonstrated. In this review, an overview
of the development of magnetism and magnetoelectric will be firstly given. The conceptual illustration and
materials used in integrated magnetoelectric will then be presented. Selections of on-chip devices from
literatures will be shown to exemplify the integrated magnetic and magnetoelectric applications. Finally,
the prospect and the direction of the future research will be discussed in the conclusion.
INDEX TERMS Magnetic, magnetoelectric, integrated devices, high frequency, sensing.
I. INTRODUCTION
Magnetic materials have a long history of usage since 4500
years ago when Chinese made the compass and also play an
important role in modern technologies like motor, electrical
generator, and transformer, etc [1]. With the development of
solid state physics, the magnetic properties and theories be-
yond the static states have been discovered and utilized includ-
ing the magnetostatic wave propagation, magnetization oscil-
lation and spintronics, which enable a wide variety of new ap-
plications especially for high frequency integrated electronics
and magnetic storage including waveguide [2], inductor [3],
filter [4], phase shifter [5], Magnetoresistive random-access
memory (MRAM) [6], spin-based transistor [7], oscillator [8]
and sensor [9] etc.
Magnetoelectrics (ME) refers to the coupling between the
magnetism and electricity and is one of the aspects of multifer-
roics which indicates the material presents two or more ferroic
properties (ferroelectricity, ferroelasticity, ferromagnetism). It
has been shown that the two-phase ME heterostructure that
couples through the mechanical strain can generate much
larger ME coupling compared with its single phase counter-
part [10]. The ME coupling effect provides a new methodol-
ogy to mediate and monitor the magnetism in the magnetic
material, and hence brings numbers of new potential features
to the devices mentioned above. The ME coupling can be
divided into two categories, namely direct ME effect (DME)
where the electric polarization is controlled by the magnetic
field and converse ME effect (CME) where the magnetization
is controlled by the electric field.
Since the emerging and prosperity of complementary
metal-oxide-semiconductor (CMOS) and thin-film technol-
ogy, both magnetic and magnetoelectric materials have been
integrated into the batch fabrication and incorporated in num-
bers of compact, power efficient on-chip devices. Among
them, the integrated magnetic material serves versatile func-
tions and will be elaborated later, DME is majorly used
for sensing applications [11] and energy harvesting [12]
where the external magnetic field is coupled to magnetic
phase and generate an electric filed at output, whereas the
CME is utilized to achieve numbers of integrated elec-
tric field tunable devices including tunable inductor [13],
filter [14], etc. The newly demonstrated nanoelectromechani-
cal system (NEMS) transmitting ME antennas [15] also rely
on the CME. A selection of on-chip radio frequency (RF)
/microwave, power and sensing devices will be shown in
this paper to elaborate the typical application structures, fea-
tures and benefits brought by the integrated magnetics and
magnetoelectrics.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/
VOLUME 1, NO. 4, OCTOBER 2021 1
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
II. MAGNETOELECTRIC EFFECT, MATERIALS AND
RELEVANT PHYSICAL MECHANISMS
As one category in multiferroic material, ME material shows
ferroelectricity and ferromagnetism simultaneously and hence
provides a potential methodology for achieving multiple func-
tions in one device and subsequently device miniaturization
[10]. The first experimental demonstration of ME effect was
achieved in single phase Cr2O3material by Dzyaloshinski
in 1960 [16], numerous efforts have been put into discov-
ering new material for larger ME coupling effect. However,
researchers found in 2000 that in respect to the transition
metal atomic orbital, the conditions forming the ferroelectric-
ity and ferromagnetism excludes each other, which explains
the difficulty of synthesize highly efficient material with large
ME effect under various temperatures [17]. On the other hand,
strain mediated two-phase ME composite, which typically
consists of an electrostricive/piezoelectric phase and a mag-
netostrictive phase, is capable to overcome the physical limi-
tation of its single-phase counterpart and present stronger ME
coupling. The two-phase ME composite utilizes the material
strain to convert the magnetic field to electric polarization in
DME case and to convert electric field to magnetization in
CME case. To quantify the strength of ME coupling, the direct
and converse ME coefficient are named and defined as [18]:
αDirect =P
H(1)
αConverse =M
E(2)
In the above equations, Pis the electric polarization, His
the applied magnetic field, Mis the magnetization and Eis the
applied electric field.
To better describe the ME coupling condition in designing
practical applications or analyzing the experimental results,
magnetically induced voltage ME coefficient is defined and
widely used [19]:
αME =E
H(3)
where His the applied magnetic field and Eis the induced
electric field.
It is clear that for two-phase ME composite material, the
intrinsic material properties and the interface condition are
of significance to ensure a decent coupling condition. As an
example, the theoretical upper limit of ME coefficient of a
L-L (longitudinally magnetized longitudinally poled) config-
uration ME composite can be given as [20]:
αME=nd33,md33,p
nsE
33 1k332+(1n)sH
33
(4)
where d33,mand d33,p are the longitudinal piezomagnetic and
piezoelectric coefficient, k33 is the electromechanical cou-
pling coefficient of the piezoelectric layer, sE
33 and sH
33 are the
elastic compliance for the piezoelectric and magnetostrictive
layers, and nis a thickness fraction of the magnetostrictive
layers.
The strong ME coupling effect can boost the performance
of DME based sensing and energy harvesting in terms of
the sensitivity or energy conversion efficiency, as well as the
function tunning range of CME based devices. In the rest
part of the section, we will present various material choices
in both piezoelectric/electrostrictive phase and magnetostric-
tive phase that are widely adopted for achieving strong ME
coupling. For the CME-based devices, since device features
are typically enabled by different magnetic film properties and
CME is adopted to alter these properties for reconfigurability,
we will discuss the magnetic properties tuning mechanisms
induced by the strain brought by CME in this section and elab-
orate the material choices and configurations in the following
sections that focus on the devices.
A. CME INDUCED MAGNETIC PROPETY TUNING
MECHANISMS
As previously mentioned, unlike DME whose applications
typically requires a strong ME coupling, CME-based appli-
cations focus more on the reconfigurability brought by the
strain-mediated magnetic properties. For a simple multiferroic
composite with a magnetic thin film deposited or glued on
a piezoelectric layer, a strain generated in piezoelectric layer
by applied electric field is transferred to magnetic thin film.
Due to the induced magnetoelastic energy and the tendency of
energy minimization, the influence of the strain upon the static
and dynamic behavior of magnetization can be quantitively
denoted by the voltage induced effective in-plane magnetic
field [21, 22], which is given as:
Hef f =3λs·Y·def f ·E
μ0Ms
(5)
where λsis the saturation magnetostriction of the magnetic
film, Yis the Young’s modulus of the magnetic film, Msis
the saturation magnetization of the magnetic film; deff is the
effective piezoelectric coefficient of the piezoelectric layer,
and Eis the electric field applied on the piezoelectric layer.
In this part, we will discuss two magnetic property tuning
mechanisms that are induced by the strain, including E-field
tuning ferromagnetic resonance (FMR) and E-field tuning
permeability. The abovementioned mechanisms serve as the
basis for the reconfigurability of multiple integrated high fre-
quency, power and sensing devices. There are some more tun-
ing mechanisms including E-field tuning magnetoresistance
and E-field tuning M-switching, the former one is utilized to
adjust the sensitivity and detection range of magnetoresistance
sensors [23], the later one is majorly adopted for magnetic
storage devices like ME random access memory (MERAM)
to control the magnetization rotation [24]. These two mech-
anisms will not be elaborated in this work considering their
rather minor applications.
1) E-FIELD TUNING FERROMAGNETIC RESONANCE
FMR refers to the resonance happens when the spontaneous
Larmor precession frequency of magnetization coincides with
the external electromagnetic wave excitation. In this case,
2VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
the external excitation signal is absorbed for magnetization
precession and attenuated due to the presence of the mag-
netic damping. FMR was first experimentally demonstrated
by Griffiths in 1946 [25]. In 1951, Kittel proposed Kittel
equation [26] that determined the FMR frequency:
fFMR
=γ(Hdc+Ha+(NyNz)4πMs)(Hdc+Ha+(NxNz)4πMs)
(6)
In equation 6, γis the gyromagnetic ratio; Hdc is the applied
DC magnetic field; Hais the spontaneous magnetic anisotropy
field, Nx,N
y, and Nzare the demagnetizing factors deter-
mined by the dimension of the magnetic material.
It Is Clearly Shown in Equation 6 With the Effective Mag-
netic Field Induced By the strain, Hdc Changes and Hence
Shift the FMR frequency.
For the magnetic film exhibiting in-plane magnetization
parallel to E-field induced effective magnetic field, the tun-
ability of the FMR frequency can be approximated as [27]:
f
f1
2
Hef f
Ha
(7)
For the magnetic film exhibiting out-of-plane magnetiza-
tion parallel to E-field induced effective magnetic field, the
tunability of the FMR frequency can be approximated as:
f
fHef f
HaMs
(8)
Fig. 1(a) presents a typical test setup for FMR absorp-
tion of magnetic material thin films. The FeGaB/PZN-PT
ME composite was placed on top of a transmission line.
The electromagnetic wave propagated in the transmission line
was coupled with the magnetic material and a notch in the
transmission coefficient S21 spectrum would appear at the
corresponding FMR frequency due to the energy absorption.
The S21 spectrum of the FeGaB/PZN-PT ME composite in the
self-biased condition (without the excitation of bias magnetic
field) is shown in Fig. 1(b) [21]. It can be noted that the
FMR where the maximum absorption peak resides can be
tuned by electric fields in the self-bias condition over a wide
range of frequency from 1.75 GHz at zero electric field to
7.57 GHz at 6 or 8 kV/cm, with a tunable frequency range of
5.82 GHz. The wide electrostatically tunable frequency range
corresponding a mean tunable frequency per unit electric field
of 15 GHz·cm·kV-1 and a tuning ratio of fmax/fmin =4.3 have
been achieved.
2) E-FIELD TUNING PERMEABILITY
Permeability is another crucial parameter of magnetic mate-
rials. High permeability materials are widely used in trans-
former and inductor applications to concentrate the magnetic
flux and hence increase the inductance and magnetic coupling
factor. With the need of the reconfigurability, the ME compos-
ite can serve as the magnetic core for these devices.
FIGURE 1. (a) Microwave absorption test setup for FeGaB/PZN-PT ME
composite. (b) Electric field dependence of the transmission coefficient
(S21) spectra of FeGaB/PZN-PT ME composite. (Reproduced from [21]).
FIGURE 2. Magnetic hysteresis loops of the FeGaB/PZN-PT multiferroic
heterostructure under different external electric fields. (Reproduced
from [21]).
The effective field generated by the strain in ME composite
changes the magnetic anisotropy field and thereafter the easy
axis direction of the magnetic material. Therefore, the effec-
tive permeability describing the magnetization response under
the applied magnetic field in a certain direction change with
the applied electric field applied on the ME composite.
To exemplify the tunning mechanism, the magnetic hystere-
sis of FeGaB/PZN-PT ME composite under different electric
fields is shown in Fig. 2 [21]. The tunable hysteresis loop
shows an in-plane anisotropy field ranging from 20 Oe at
0 kV/cm and 700 Oe at 6 kV/cm. The different slopes of
VOLUME 1, NO. 4, OCTOBER 2021 3
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
TAB L E 1. Comparison of Commonly Used Ferro/Piezoelectric/Electrostrictive Materials for ME Applications
the hysteresis loop indicate the applied electric field have
changed the magnetization response and hence the effective
permeability in in-plane direction.
To quantify the effective permeability μeff with the presence
of strain, the following equation is given in [13]:
μef f =4πMs
Hef f
+1(9)
where the effective magnetic anisotropy field can be given as:
Hef f =Ha+Heff (10)
where Hais defined in equation 6 and Heff is defined in (5)
as the electric field induced effective anisotropy magnetic field
in the magnetic films.
B. PIEZOELECTRIC/ELECTROSTRICTIVE MATERIALS FOR
MAGNETOELECTRICS
Among piezoelectric/electrostrictive materials, lead zirconate
titanate (PZT), lead magnesium niobate–lead titanate (PMN-
PT), lead zinc niobate–lead titanate (PZN-PT) and aluminum
nitride (AlN) have been widely used in integrated ME de-
vices to convert electric excitation to strain. Table 1 shows
a cross comparison of the most commonly used piezoelec-
tric/electrostrictive materials for integrated ME applications.
PZT is a polycrystalline ceramic ferroelectric material with a
high electromechanical coupling coefficient and an effective
piezoelectric coefficient of 200 to 300 pC/N [28] at mor-
photropic phase boundary (MPB) composition corresponding
to Zr/Ti ratios of about 58/42. The PZT piezoelectric ceramic
also has the advantages of low-cost and easier shaping process
[29]. The main drawback of PZT materials is that the contam-
ination during the process is often unavoidable due to the lead
in it, which greatly degrades its application in magnetoelectric
devices [30]. Moreover, the relatively high acoustic loss of the
PZT film at high frequency range prevents it from achieving
an efficient ME coupling. In Table 1, PZT materials are clas-
sified into hard and soft PZTs, where the hard PZT materials
usually have smaller piezoelectric coefficient with higher en-
ergy efficiency. Depending on the circumstance of the specific
applications, various types of PZT materials are commercially
available. Different from the PZT film, aluminum nitride
(AlN) is a textured polycrystalline non-ferroelectric material
with an advantage of lower loss tangent and compatibility
with CMOS technology. Another unique advantage of AlN
is its high electric Curie temperature, which is desirable for
electronics applications working under high temperature of
inferior heat dissipation environment. However, the moderate
piezoelectric coefficient of about 2 pC/N makes it unsuitable
for ME devices requiring high ME coupling [31]–[33].
Compared with PZT and AlN, lead magnesium niobate–
lead titanate (PMN-PT) and lead zinc niobate–lead titanate
(PZN-PT), as single crystal ferroelectric materials, show
larger piezoelectric coefficients and lower loss tangents. The
piezoelectric performance of these materials highly depends
on their composition and crystal orientation. The composition
here means the element ratio of rhombohedral phase (PZN
or PMN) and tetragonal phase (PT). The PZN-PT composi-
tion that induce MPB with (001) crystal cut can give a d33
piezoelectric coefficient as high as 2800 pC/N [34]. For the
PMN-PT film, a 31% PT and optimized cut give a d31 of about
-1800 pC/N and a d33 of about 2000 pC/N [35]. Although
the forementioned two materials deliver superior piezoelectric
performance, the difficulties of composition control during the
deposition of these materials may induce the lack of perfor-
mance repeatability. The grain growth of these single crystal
materials is also challenging. Two major deposition methods
include flux method and Bridgman method [36].
C. PIEZOMAGNETIC/MAGNETOSTRCTIVE MATERIALS FOR
MAGNETOELECTRICS
In general, the desired properties of piezomag-
netic/magnetostrictive phase in ME composite for achieving
a strong ME coupling are high saturation magnetostriction λs
and high piezomagnetic coefficient dλ/dH. Depending on the
specific applications, other magnetic and elastic properties
of the materials may also be considered. The low magnetic
loss tangent, low FMR bandwidth and low coercive field
are essential for high frequency application. Some particular
properties like E effect where the Young’s modules of
materials vary with the applied magnetic field, are also
needed for specific applications including sensors.
A cross-comparison of commonly used piezomag-
netic/magnetostrictive materials for ME composite is given
in Table 2 including Terfenol-D (Tb0.7Dy0.3Fe2), Galfenol
(Fe81Ga19 ), FeGa, FeCoSiB, FeGaB, FeBSiC (Metglas) and
FeGaC.
Based on the Terfenol-D with the largest magnetostriction,
the Terfenol-D/PZN-PT composite with giant electric field
induces effective magnetic anisotropy field of 3500 Oe, cor-
responding to a magnetoelectric coefficient of 580 Oe·cm/
kV [44]. The magnetoelectric coefficient is almost 4 times
4VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
TAB L E 2. Comparison of Commonly Used Piezomagnetic/Magnetostrictive Materials for ME Applications
FIGURE 3. (a) Boron content dependent magnetostriction property of FeGaB thin film material. (b) Boron content dependent FMR linewidth of FeGaB
thin film material. (Reproduced from [48]).
of that in Fe3O4/PZN-PT [22]. Even though Terfenol-D ex-
hibits giant saturation magnetostriction of 1600 ppm, it is
lossy and has a high saturation field of several kOe, which
induces a relatively small piezomagnetic coefficient. Com-
pared with Terfenol-D, binary FeGa alloys show a high satura-
tion magnetostriction constant of 400 ppm for single crystals
(Galfenol(Fe81Ga19 )) and 275 ppm for directional solidified
polycrystalline alloys (FeGa), a low saturation field of about
100 Oe and a high saturation magnetization of 18 kG. How-
ever, the FeGa films have a large FMR linewidth of 450-
600 Oe at X band and strong loss at microwave frequencies
[45]. The shortcoming indicates that the FeGa alloys are far
from being an ideal magnetic material to be incorporated into
RF/microwave magnetoelectric devices. Boron (B), as a well-
known metalloid element, has been added into the FeGa and
FeCo films to improve their magnetic properties. It has been
experimentally demonstrated that the addition of B in FeCo
films contributes to an excellent soft magnetization and mi-
crowave properties by refining the grain size and diminishing
magnetocrystalline anisotropy [46,47]. Moreover, the addition
of Boron in the composition of FeCoSiB and FeGaB thin films
contributes to the demonstration of giant ME coefficients and
low loss tangents in ME devices RF/microwave devices.
FeGaB material exhibits excellent performance on mag-
netic softness, saturation magnetization, saturation magne-
tostriction, piezomagnetic coefficient, E effect, magnetome-
chanical coupling factor and has been regarded as the most
ideal piezomagnetic film material for ME devices [48]. As
indicated in [48], with the addition of B rising from 0% to
21% in FeGa, the coercivity Hcdecreased from about 100 Oe
to 1 Oe and the effective anisotropy field Hkdropped from
120 Oe to 15 Oe due to the material phase transition (MPB).
The reduced coercivity and anisotropy significantly enhances
magnetic softness. Additionally, as shown in Fig. 3(a), the B
content induced phase transition results in a lower threshold
magnetic field to generate a magnetostrictive response. The
maximum magnetostriction of 70 ppm achieved at 12% B
content was three times of that of the binary FeGa films.
The FMR linewidth of FeGaB films also dropped dramatically
from 700 Oe for binary FeGa films to 16 Oe at 21% B content.
The maximum piezomagnetic coefficient of the FeGaB films
has been reported as about 7 ppm Oe-1, which is much higher
than multiple magnetostrictive materials,including Terfenol-D
(Tb-Dy-Fe), Galfenol (Fe-Ga) and Metglas (FeBSiC). The
comparison was made in [21] and graphically shown in Fig. 4.
The maximum piezomagnetic coefficient can be further en-
hanced to 12 ppm Oe-1 after annealing due to the effective
release of residual stress inside the film and the reduction of
total anisotropy energy [49].
III. INTEGRATED MAGNETIC AND MAGN ETOELECTRIC
FOR POWER, RF, AND MICROWAVE APPLICATIONS
Integrated magnetic materials play a significant role in
enabling high performance or unique features in multiple
power, RF and microwave devices. The static and dynamic
properties of these materials are widely utilized to achieve
high efficiency magnetic flux guide, frequency selectivity
VOLUME 1, NO. 4, OCTOBER 2021 5
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
TAB L E 3. Integrated Magnetic and Magnetoelectric Devices for Power, RF, and Microwave Applications
FIGURE 4. Piezomagnetic coefficients (dλ/dH) of different types of
magnetostrictive alloys. (Reproduced from [21]).
by spontaneous resonance, magnetostatic wave guide, spin
controlling, etc. As previously mentioned, by introducing
strain-mediated magnetoelectric composites, the magnetic
properties of these material can be altered by applying electric
field, which provides possibilities for novel electrostatically
highly energy-efficient, compact, miniature and lightweight
tunable devices for power, RF and microwave electronics. In
this section, several integrated magnetic and magnetoelectric
devices including high performance transformer, inductor
with laminated magnetic core, voltage tunable inductor, E-
and H- field tunable bandpass filter and NEMS ME antennas
are presented. The selected applications are good representa-
tives for power components, basic circuit components, RF and
microwave signal processing block and RF and microwave
signal transmitting/ receiving block, respectively. A summary
of the applications presented in this section is given in Table 3.
A. ON-CHIIP TRANSFORMER, INDUCTOR WITH
INTEGRATED MAGNETIC CORE AND VOLTAGE TUNABLE
INDUCTOR WITH ME COMPOSITE
The inductor and transformer are two types of devices whose
operations rely on the magnetic flux variation and energy
conversion. As the rapid development of integrated RF, mi-
crowave and power circuits, the high Q, compact induc-
tor with sufficiently large inductance and transformers with
high power efficiency and coupling coefficient are highly de-
manded. The integrated magnetic materials, serve as the mag-
netic core, can boost the inductance and the transformer cou-
pling condition by concentrating the magnetic flux generated
during the operation. The mostly used magnetic materials for
core are always high-permeability materials including permal-
loy (NiFe) [55], NiFeW [56], NiFeZn [57], NiZn ferrite [58],
6VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
FIGURE 5. (a) Optical image of the on-chip inductor with solenoid structure. (b) Inductance measurement result of inductors with magnetic core and air
core. (c) Quality factor measurement results of inductors with magnetic core and air core. (Reproduced from [3]).
FeCoSiB [59], FeCoCu [60], FeGaB [3], [61], etc. However,
there are two limitations brought by the integrated magnetic
core: 1). The relatively thick magnetic core in on-chip de-
vices induces large eddy current loss during high frequency
operation, which limits the quality factor; 2). The FMR of the
magnetic core causes energy absorption in the corresponding
frequency range, which diminishes the power efficiency and
working frequency range. To mitigate the high eddy current
loss issue, the integrated magnetic core was engineered as a
laminated structure [61]–[63] with thinner magnetic material
layers and isolation spacer in between. The thinner magnetic
material has higher equivalent resistance and hence reduce the
eddy current induced by the time-varying magnetic field. As
for the loss caused by FMR, the Kittel [64] equation sug-
gests that by determining proper magnetic film dimensions,
the demagnetizing factor can be manipulated such that the
FMR frequency is higher, which subsequently expand the op-
erational frequency range [63]. The most commonly adopted
inductor and transformer coil structures include solenoid and
spiral/planner structure, where the solenoid structure [55],
[58], [61], [63] incorporates the magnetic core between the
wires at upper and lower layers whereas in the spiral/ planer
[65]–[68] structure the magnetic core layer is deposited be-
neath or between the planer wires to form the magnetic
circuit.
In 2014, Gao [3] reported a series of integrated induc-
tors using multilayer FeGaB (100 nm)/Al2O3(5 nm)/FeGaB
(100 nm) sandwich structure as the magnetic core. The au-
thors adopted solenoid structure for the inductor design and
the optical image of the as-fabricated device is shown in
Fig. 5(a). As shown in Fig. 5(b), compared with the air core
inductor with same dimensions, the magnetic core boosted the
inductance by nearly 100% in the frequency range of up to 3
GHz. The inductance showed a flat characteristic with a peak
inductance of over 3 nH, which indicated that the laminated
magnetic core was of help to eliminate the eddy current loss
in high frequency range and the energy loss contributed by
FMR. The quality factor in the frequency range of up to 3
GHz is shown in Fig. 5(c), where the peak quality factor of
the inductor with magnetic core was over 10 and showed a
better performance compared to the air core inductor in the
relatively low frequency.
Mullenix [62] reported an on-chip micro-transformer with
integrated laminating magnetic core in 2013. The device turns
ratio is 1:1. As shown in the Fig. 6(a), the transformer has a
dual-coils solenoid structure. The authors fabricated a series
of devices with 8 turns coils, 16 turns coils, 32 turns coils,
and 32 turns coils with air core for comparison. The cross-
sectional view of the coil is shown in Fig. 6(b), a 2.8 µm thick
NiFe/AlN laminated film was adopted as magnetic core and
the optimized thickness of the insulating AlN layer was deter-
mined to be 7 nm. The inductances of a pair of inductors with
a 500 µm×5.6 µm core connected in series with difference
turn numbers are shown in Fig. 6(c). The 32 turns AC in the
plot indicates the measurement result of air core inductor. The
introducing of the magnetic core boosted the inductance by
60 times, namely 565 nH in the operating frequency range.
Fig. 6(d) shows the maximum Q factor achieved by 32 turns
magnetic core transformer was 6.3 and the maximum coupling
factor is 0.97, indicating a superior coupling between the
primary and secondary coil. The results show the advantage
of adopting laminating magnetic core since the pure magnetic
material core with same thickness has lower roll-off frequency
than the designed operation range.
Tunable RF inductors provide another possible solution for
tackling the trade-off between inductance, quality factor and
operating frequency range, since the external electric signal
can lead to the control and improvement of these 3 parameters
at the same time. There have been two main branches of con-
trol strategies for tunable inductors in microelectromechani-
cal systems (MEMS). One mechanism is to adjust the turns
or numbers of inductor coils. Based on this, Park changed
the turns or length of inductor coils by MEMS switch and
achieved an operation frequency of 2.4 GHz, a Q-factor of
3 and extremely large tunability of 187.5% [69]. The main
drawback of this MEMS switch control strategy is the discrete
tunability of inductance limited by the number of switches
[69, 70]. In order to overcome this shortcoming, the inductor
coupling control method with a secondary coil was put for-
ward and applied to the fabrication of related devices, with
VOLUME 1, NO. 4, OCTOBER 2021 7
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
FIGURE 6. (a) Optical image of the on-chip transformer. (b) Cross-sectional view of the coil. Measured (c) inductance, (d) quality factor and coupling
coefficient for the devices with 500 μm X 5.6 μm core with 8, 16, and 32 turns. The “32 Turns AC” refers to the air core transformer of identical geometry
with 32 turns. (Reproduced from [62]).
an operating frequency ranging from 4.35 GHz to 7 GHz,
a quality factor of 10 and an inductance tunability from
30% to 90% [71], [72]. However, the related work was
greatly limited by the complicated structure and fabrication
procedure. Another perspective is to change the properties
of the magnetic core or the magnetic field distribution in
it. Following this direction, researchers tried to control the
magnetic flux by thermal actuation [73] or metal shield [74].
DC current was also widely used to modify the properties of
magnetic core. For example, Salvia realizes a 40% increase
in inductance, an inductance tunability of 15% and a Q-factor
from 5 to 11 in the tunable on-chip inductors operating up to
a frequency of 5 GHz using patterned permalloy laminations
[67]. Unfortunately, both methods exhibit limited inductance
tunability, high power consumption, significant thermal effect
and excessive noise [67], [75]–[77].
With the development of magnetoelectric materials, re-
searchers tended to use ME coupling to adjust the perme-
ability of magnetic core and tune the inductance, aimed at
improving the energy efficiency, reducing joule heating and
increasing inductance tunability. The CME tuning mechanism
is elaborated in Section II. In order to ensure higher induc-
tance density and efficient utilization of magnetic film, the
most popular spiral and solenoid structures have been widely
used in electrostatically tunable inductor. A solenoid type
magnetoelectric inductor fabricated with a layered multifer-
roic composite core was proposed by Lou, which consisted of
two layers of Metglas magnetic ribbons and one PZT piezo-
electric slab [78]. The inductance showed a strong electric
field tunability, with maximum inductance change of 450%,
250%, and 50% for operation frequencies of 1 kHz, 100 kHz,
and 5 MHz, respectively. The estimated energy consumption
to achieve maximum tuning 450% is only about 1.3 mJ, in-
dicating that the fabricated tunable magnetoelectric inductors
were essentially passive and energy efficient and achieved a
giant tunability. Due to the decreased permeability, increased
skin depth and reduced core eddy current loss, the quality fac-
tor is improved from about 3 at zero electric field to about 8.5
at an external field of 12 kV/cm. The giant tunability achieved
in such magnetoelectric inductors with low power consump-
tion makes them ideal candidates for novel power efficient and
compact electronic devices. In order to extend the operation
frequency range and enhance the quality factor by reduc-
ing the eddy current loss, Lin fabricated PZT/Metglas/PZT
voltage tunable inductors with single layer Metglas film and
multiple piezoelectric slabs [79]. An 80% increase in Q-factor,
10 times of extension in operational frequency and 3 times of
enhanced quality factor were observed. In addition, a 150%
8VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
FIGURE 7. Integrated gigahertz FeGaB/Al2O3/PMN-PT voltage tunable ME inductor. (a) Optical image and structure model of integrated voltage tunable
inductor with 3.5 turn and a magnetic film of 340 μm X 800 μm. (b) Measured inductance and (c) Q-factor of integrated FeGaB/PMN-PT tunable inductor
with applied electric field from 0 to 10 kV/cm. (Reproduced from [13]).
improvement in quality factor and 100 times of extension in
operational frequency could be found in ME inductors with
a single PZT slab, whereas a 235% improvement in quality
factor and 200 times of extension in operational frequency
were realized in its double PZT slabs counterpart.
Though the PZT/Metglas/PZT ME inductors show a high
performance in quality factor, inductance density and a gi-
ant tunability of 450% in the kilohertz frequency band, the
Metglas/PZT multiferroic heterostructures are not applicable
for the gigahertz applications due to their low FMR fre-
quency. In addition, these inductors fail to realize compati-
ble integration on CMOS circuits, which prevents their ap-
plications in RF/microwave electronics. To realize gigahertz
range frequency operation, Chen has recently demonstrated a
voltage tunable GHz integrated magnetic inductor by using
the FeGaB/Al2O3/PMN-PT magnetoelectric composites, as
shown in Fig. 7(a) [13]. Compared with materials like NiZn,
the high permeability further improved by annealing, FMR
frequency of 1.85 GHz with a narrow FMR bandwidth of
1620 Oe, large magnetostriction of FeGaB films lead to high
inductance density, operation band in GHz range, relatively
low magnetic loss and high frequency tunability, respectively.
The inductor showed a relatively constant inductance and
quality factor value in the frequency range of 0.5 to 2 GHz.
The alumina spacer reduced the thickness of single-layer mag-
netic material, thereby increasing the sheet resistance and re-
ducing the eddy current loss. The E-field tunability test results
shown in fig. 7 (b) and (c) indicate that in this frequency range,
with an applied electric field up to 10 kV/cm, the inductance
rises from 1.2 to 3.5 nH and the quality factor rises from
2.7 to 5.2. The device achieved a maximum inductance contin-
uous tunability of 191% at 1.5 GHz, which was significantly
larger than other integrated inductors tuned by magnetic flux,
DC bias current and coil coupling.
B. INTEGRATED RF AND MICROWAVE FILTERS AND
VOLTAGE TUNABLE FILTERS WITH ME COMPOSITES
Filter is an essential component in RF/microwave signal pro-
cessing circuits and blocks. There are several types of filters
where the integrated magnetic and magnetoelectric materials
function as a core part. In a typical filter based on lumped
element [80,81], the magnetic materials serve as the core of
inductors in the filter topology and the function is similar as
forementioned inductor components. FMR absorption and
magnetostatic wave manipulation properties of magnetic
materials are also widely utilized for achieving band-stop and
band-pass characteristics. Band-stop filters utilizing FMR
absorption typically consist of a transmission line that couples
with the magnetic materials. When the frequency of electro-
magnetic wave propagating in the transmission line coincides
with FMR frequency of the material, a maximum attenuation
and hence a stop band is generated. Multiple materials includ-
ing transition metal [82], ferrite [83]–[85] and soft magnetic
composite [4] are adopted for these band-stop filters. Mag-
netostatic wave manipulation property in ferrite is utilized to
achieve non-reciprocal band-pass characteristic [14], which
VOLUME 1, NO. 4, OCTOBER 2021 9
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
FIGURE 8. Integrated NiZn/PMN-PT dual E- and H-field tunable ME band-pass filter. (a) Schematics of integrated tunable inductor. (b) S12 and S21 under
abiasmagneticfieldof400Oe.(c)S
11 under different bias electric fields. (d) S21 under different bias magnetic fields. (Reproduced from [14]).
will be elaborated later in this part. The third type of integrated
filter utilizes the mechanical resonance to realize frequency
selectivity [86], where the input excitation signal is trans-
formed into acoustic wave at the input resonator and propagate
to the output resonator thorough the released material layer.
Reconfigurability is a crucial feature in filter design consid-
ering the need for wide band operation, the size of the signal
processing block and the yielding cost. MEMS switches [87]–
[89] are adopted to realize the reconfigurability. However,
this mechanism suffers from the tuning discontinuity which
limits its flexibility. The integrated filters that incorporate the
magnetic materials mentioned above are usually magnetically
tunable. In [4], He presented an integrated tunable band-stop
filter that achieved a tunability of 70% under a low magnetic
field of 400 Oe working in C-band. However, for some higher
frequency filter applications, the relatively lossy ferrite has to
be adopted and a large magnetic field is required for tunabil-
ity [83]. Meanwhile, the instrument for applying a magnetic
field is usually bulky and power consuming, which makes the
magnetically tunable filters an inferior choice. Integrated ME
materials offer a possibility to achieve on-chip electric field
tunable filter devices. The advantages of ME tuning mech-
anism are continuous tuning, low power consumption and
easily-defined tuning effect range (determined by the size of
ME materials). Current efforts are still committed to mitigate
the magnetic material performance degradation induced by
the deposition or bonding of the piezoelectric phase in ME
materials.
In 2015, the first integrated dual E- and H- field tunable
band-pass filter was demonstrated in an inverted-L-shape mi-
crostrip structure with a multiferroic heterostructure formed
by a spin-sprayed NiZn ferrite film and a PMN-PT piezoelec-
tric slab, as shown in Fig. 8(a) [14]. The signal from the input
port is transmitted as magnetostatic surface wave (MSSW).
The forward and backward wave transmissions happen on
bottom and top surface of the ferrite film. The passband fre-
quency of the filter is determined by the dispersion relation of
the MSSW. Due to the discrepancy of radiation resistance on
the top and bottom surfaces of the ferrite film, a non-reciprocal
transmission characteristic can be observed [90], [91]. By ide-
ally rotating the NiZn ferrite film by 45 degrees, the standing
wave resonance and the multi-passband are diminished since
the reflected wave at the edge of ferrite film is converted to
magnetostatic back volume wave (MSBVW) which will decay
fast at the filter operating frequency. The detail of the filter
operation principle is presented in [92]. As shown in Fig. 8(b),
the 15.8 dB difference of S12 and S21 in band shows the
non-reciprocal transmission. The central frequency could be
electrically tuned from 2.075 GHz to 2.295 GHz when the
out-of-plane electric field rises from 0 kV/cm to 4 kV/cm,
as shown in Fig. 8(c), namely an E-field frequency tunability
of 55 MHz/(kV/cm). In addition, a tunable central frequency
from 3.78 GHz to 5.27 GHz could be realized by increasing
the DC magnetic field from 100 Oe to 400 Oe, corresponding
to an H-field frequency tunability of about 5 MHz/Oe, as
shown in Fig. 8(d).
10 VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
FIGURE 9. Integrated FeGaB/Al2O3/AlN dual E- and H-field tunable ME band-pass filter. (a) Schematics of integrated tunable filter. (b) S11 and S21 of the
tunable filter at zero bias field. (c) Measured resonant frequency as a function of DC magnetic field. (d) Measured resonant frequency as a function of DC
voltage across the thickness direction of the AlN film. (Reproduced from [86]).
In 2016, an integrated RF tunable bandpass filter based on
two coupled elliptic-shape nano-mechanical resonators with
FeGaB/Al2O3/AlN multiferroic heterostructure on a bottom
Pt electrode were reported [86] and schematically shown in
Fig. 9(a). The filter relies on the acoustic wave propagation
to achieve signal transmission between the two ports, the
two resonators work in the contour mode whose resonance
frequency is determined by the width of the coupled structure.
The filter exhibited a return loss of -11.2 dB, an insertion loss
of 3.4 dB and a quality factor of 252 at the central frequency
of 93.165 MHz, as shown in Fig. 9(b). The tunability test
showed a H-field frequency tunability of 5 kHz/Oe and an
E-field frequency tunability of 2.3 kHz/V, presented in fig.
9(c), (d), respectively. The tunability is realized based on the
E effect, where the Young’s modulus of FeGaB changes
with applied fields.
C. STATE- OF-ART INTEGRATED ME ANTENNAS
Conventional antennas rely on the direct excitation of current
or voltage to control the motion of electrons inside resonators
for radiation and typically have sizes ranging from one-tenth
to one-half of the electromagnetic (EM) wavelength λ0[93],
[94]. Multiple miniaturization techniques have been applied
to design miniature antennas with different structures, such as
dipoles, monopoles, slot, Z-type, metamaterial loaded anten-
nas [95]. Electrically small antenna performance is restricted
by the Chu–Harrington limit [96]–[99]. The inherent high
quality factor (Q-factor) of these antennas limits both the
radiation efficiency and operating bandwidth.
The concept of ME antennas was first put forward in [100,
101]. Unlike the conventional antennas that use oscillating
charges to induce EM waves, the ME antennas use magnetic
dipole moment oscillations that are acoustically actuated at
their electromechanical resonance (EMR) rather than the EM
wave resonance. Therefore, though there is no proof that the
Chu–Harrington limit has been overcome, the antenna di-
mensions of the mechanical antennas can be decoupled with
the wavelength of EM waves. Since the velocity of acoustic
waves is orders of magnitude slower than EM waves at the
same operation frequency, the application of ME antennas
will lead to 1–2 orders of miniaturization in antenna dimen-
sion. In addition, ME mechanical antennas based on novel
ME composites could maintain a high radiation efficiency,
wide bandwidth, low power loss and even high transmission
rates and adjustable multiband under appropriate design and
nonlinear modulation strategy, which may become an excel-
lent solution to very low frequency (VLF) communication
[102], broadband wearable and implantable biomedical de-
vices [103]–[106] and multiband portable devices [107].
The ME antennas consisting of one layer of piezoelectric
material and one layer of magnetostrictive material rely on the
bulk acoustic wave (BAW) resonator to transfer the dynamic
VOLUME 1, NO. 4, OCTOBER 2021 11
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
FIGUR E 10. (a) Illustration and explanation of the new antenna
mechanism. (b) Illustration and explanation of the ground plane effect.
(Reproduced from [108]).
strain across different layers. The illustration of the novel ME
antenna mechanism is shown in Fig. 10(a) [108]. On the trans-
mitting side, by applying RF electric field, the mechanical
resonance would induce an alternating strain wave/acoustic
wave that can be directly transferred to the upper ferromag-
netic thin film. The acoustic wave would then induce a dy-
namic change of the magnetization due to the strong piezo-
magnetic effect and generate a magnetic current for radia-
tion. Reciprocally, on the receiving side, the RF magnetic
field component of the electromagnetic wave can induce an
acoustic wave on the ferromagnetic layer. The acoustic wave
is then transferred to the piezoelectric layer and converted
to the electric voltage output. Different from the cancelling
effect of imaging currents in the ground plane of radiation in
conventional EM antennas, the usage of magnetic currents for
radiation in ME antennas would provide a 3 dB gain enhance-
ment due to in-phase image currents attaching on the ground,
as shown in Fig. 10(b). This ground plane immunity property
can provide a variety of applications on the metallic surface
and the human body which is also considered as a ground
plane.
The first batch of ME antennas were fabricated by Nan in
2017, based on laterally-vibrating nano plate resonator (NPR)
and vertically-vibrating film bulk acoustic resonator (FBAR)
[15]. The ME Antenna with NPR consisted of a rectangu-
lar resonating plate with a single-finger bottom Pt electrode
and a thin film FeGaB/AlN active resonant heterostructure,
as shown in Fig. 11(a). As shown in Fig. 11(c), a high ME
coupling coefficient of αME =6kVOe
1cm1can be ob-
tained at EMR frequency fr,NPR of 60.68 MHz without DC
bias magnetic field, indicating a strong ME coupling. The
strong ME coupling leads to a high electromechanical trans-
duction efficiency, low loss and effective interaction between
acoustic resonance and RF magnetic field component of EM
waves, which contributes to a high-quality factor Q of 930
and an electromechanical coupling coefficient (kt2) of 1.35%,
as shown in Fig. 11(b). The ME NPR antenna achieves 1–2
orders of magnitude miniaturization over state-of-the-art com-
pact antennas without performance degradation.
The ME FBAR antenna was composed of a suspended Fe-
GaB/AlN ME circular disk, as shown in Fig. 11(d). Different
from ME NPR antennas, the ME FBAR antenna operated
at GHz in thickness extensional vibration mode, as shown
in Fig. 11(e). The ME FBAR antennas exhibited a peak re-
turn loss of 10.26 dB, a Q-factor of 632 and a mechanical
resonance frequency of 2.53 GHz. The operating frequency
(same as the mechanical resonance frequency) f0,FBAR can be
changed by adjusting the thickness Tof the ME heterostruc-
ture due to the relation f0,FBAR 1
TEeq
ρeq , where Trepresents
the thickness of the FBAR resonator; Eeq is the equivalent
Young’s modulus; ρeq is the equivalent density of the ME
composite. The FBAR antenna achieved an antenna gain of
-18 dBi at the resonance frequency of 2.53 GHz based on the
measured radiated signal (S12) and received signal (S21), as
shown in Fig. 11(f). Compared with the state-of-art conven-
tional compact antennas operating at the same frequency, the
dimension of the ME antenna is 1 or 2 orders smaller due to
the guided acoustic wave with much lower velocity compared
to the EM wave.
The designed and fabricated NPR and FBAR ME antennas
could realize different modes of vibration and radiation at
both very high frequency (VHF, 60 MHz) and ultra-high fre-
quency (UHF, 2.525 GHz) operation frequency bands. More-
over, the similar microfabrication process of both NPR and
FBAR based antennas on the same silicon wafer allows the
integration of broadband ME antenna arrays working in the
frequency range from tens of MHz (NPR with large lateral
dimensions) to tens of GHz (FBAR with thinner AlN thick-
ness) on one chip. These ultra-compact ME antennas serve
as the potential candidates for future communication systems,
internet of things, wearable antennas, bio-implantable and
bio-injectable antennas, smart phones and wireless commu-
nication systems.
IV. I NTEGRATED MAGNETIC AND MAGNETOELECTRIC
FOR SENSING APPLICATIONS
During the past decades, the ultra-sensitive magnetic field
sensors were highly demanded and utilized in various field
including biomedical applications, geographic detection,
electric malfunction diagnosis and information technologies,
etc. For practical usage, more figures of merit should be
considered, including working bandwidth, temperature
stability, linearity, power consumption and cost [109].
Currently the best limit of detection (LOD) of single digit
fT/Hz1/2 at 1 Hz was achieved by superconducting quantum
interference device (SQUID) magnetometer [110]. However,
the SQUID requires near-zero working temperature and hence
bulky and expensive. Integrated magnetic and magnetoelectric
sensors, which take the advantage of its silicon/CMOS
compatibility and size suitable for large sensor arrays, have
drawn much interest in recent years and notable progresses
have been made. The magnetic or magnetoelectric materials
integrated in these sensor applications typically have electric
or magnetic properties that are sensitive to the external weak
12 VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
FIGUR E 11. Magnetoelectric (ME) nanoplate resonator (NPR) and film bulk acoustic resonator (FBAR) antennas with gigantic ME coupling. ME NPR
antenna: (a) Scanning electron microscopy (SEM) images of the fabricated ME NPR antenna. (b) Admittance curve and Butterworth–van Dyke model
fitting of the ME NPR. The inset shows the schematic of the cross-section of the ME heterostructure. (c) ME coupling coefficient (left axis) and the
induced ME voltage (right axis) versus the frequency of RF magnetic field excitation HRF. FBAR antenna: (d) SEM photo. (e) return loss S22 and (f)
radiating characteristic (S12) and receiving characteristic (S21 ) at resonance of the FBAR device. (Reproduced from [15]).
TAB L E 4. Integrated Magnetic and Magnetoelectric Devices for Magnetic Field Sensing Applications
magnetic field, which directly induces a measurable electric
output in thin film-based devices, or changes the mechanical
properties of the sensor structure in MEMS devices and
subsequently the electric output. The prevalent integrated
magnetic sensors include ME sensor [11], [111]–[113],
magnetoresistance (MR) sensor [114], [115], permanent
magnet-based sensor [116]–[120], and fluxgate sensor [121].
In this section, the principle and typical sensor structure
of first three types of sensors will be elaborated where the
integrated magnetic or magnetoelectric material serves as
the sensing element. In the fluxgate sensors, the magnetic
material serves as core which shares the similar function as
integrated inductors and transformers. A cross comparison of
the sensors discussed in this work is given in Table 4.
A. INTEGRATED ME SENSOR
The two-phase ME composite that consists of a piezoelec-
tric/electrostrictive phase and a magnetostrictive phase is ca-
pable to directly convert the magnetic moment to the electric
polarization and subsequently the output voltage through the
mechanical coupling between the two phases, which enables
the possibility of its sensing application. The sensing perfor-
mance of the composite is majorly determined by 1). The ma-
terial properties of both phases including the piezoelectric and
magnetostrictive constant, stiffness, dielectric constant and
permeability. 2). The mechanical structure issues including
the volume and thickness ratio of two phases, the mechanical
coupling condition. 3). The operating mode which determines
the mechanical deformation and polarization directions [122].
VOLUME 1, NO. 4, OCTOBER 2021 13
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
Earlier works [123], [124] on ME sensors focused on form-
ing the bulk ME laminate structure by epoxy bonding to
detect the external weak magnetic field thorough direct ME
coupling, where a strain induced by the magnetic field in mag-
netostrictive phase is transferred to piezoelectric phase and
reflected at the electric output. However, the major drawback
of this scheme is the narrow detectable bandwidth considering
the mechanical resonance characteristic of the sensor and the
high flicker noise in the low frequency range. A DC bias field
is also needed for optimizing the ME coupling coefficient
in the measurement frequency range. Moreover, the epoxy
glued ME sensors suffer from the temperature stability and
device fatigue issues. Some efforts have been put into optimiz-
ing the sensor configuration and structure to achieve higher
sensitivity by optimizing sensor geometry [125], improv-
ing the lamination process [126], performing a differential
sensor configuration [127] and improving the output signal
processing [128].
To tackle the above-mentioned bonding issues, thin-film
technologies are adopted to deposit the two phases in the ME
composite, which enables a more desirable bonding condition
and hence improves the ME coupling coefficient. The mostly
used soft magnetic materials for magnetostrictive phase in-
clude FeGa [129], FeCoSiB [130] and FeGaB [131], while
the piezoelectric phase is typically AlN [132], PZT [129]
and PVDF [133]. The frequency conversion technique was
first proposed by Jahns [134] and was widely adopted [135],
[136] to expand the operating frequency range of the ME
sensor, particularly in the low frequency regime. The tech-
nique utilizes the non-linear magnetostrictive response and
converts the low frequency magnetic field component to a
frequency-mixing component close to the resonance by ap-
plying a modulation magnetic field at resonance frequency.
The frequency mixing component intensity was utilized to
determine the weak magnetic field under detect.
Integrated ME sensors has its unique advantage in form-
ing the well-arranged sensor array, compatibility of CMOS
process and power efficiency. Currently, the principle of in-
tegrated ME sensor includes E effect and the direct ME
coupling. Like the bulk ME sensors, the integrated ME sensor
that relies on the direct ME coupling can only detect the AC
magnetic field with frequency close to its resonance. An ex-
ternal DC bias magnetic field is also required to optimize the
ME coupling and LOD. The E integrated ME sensor, on the
other hand, utilizes the E effect in magnetostrictive phase,
where the Young’s modulus changes with the magnetic field
under detect. The mechanical characteristic of the sensor is
then altered and can be detected by optical detection of the de-
flection [111], close-loop measurement of the resonance fre-
quency [112], open-loop measurement for sensor admittance
[131] and direct output voltage readout through a directional
coupler [113]. The E ME sensor provides the possibility
for wideband operation and low frequency/DC magnetic field
detection. The two types of integrated sensors are exemplified
below in detail to illustrate the device structures and working
principles.
In [11], Su proposed to use AlScN/FeCoSiB laminate ME
structure for sensing application based on direct ME coupling.
The simplified sensor structure is shown in Fig. 12(a), where a
1µm thick AlScN layer and a 2 µm thick FeCoSiB layer were
deposited on a poly-silicon cantilever. The Mo top electrode
and the Ti/Pt bottom electrode are not shown in the schematic.
The polysilicon cantilever structure was obtained through a
TMAH etching process. The as-fabricated sensor was pack-
aged and mounted on the test board shown in Fig. 12(b) with
a wafer-level transient-liquid-phase process. A series of AlN
based sensors were also fabricated for performance compar-
ison. Fig. 12(c) demonstrates the ME coefficient measure-
ment results of sensors with different piezoelectric phase. The
AlScN material provides a factor of 2 improvement on the ME
coefficient compared to its AlN counterpart, reached a value
of 1580 V/cm·Oe at 8185 Hz. The electric noise spectrums
of both sensors are given in Fig. 12(d) with the reference
audio noise. The peak position that appears in the spectrum
coincides with that of ME coefficient. The LOD of each sensor
was calculated by finding the ratio of the ME coefficient to the
electric noise density at resonance. The best LOD achieved
was 55 pT/Hz1/2 . It was also shown that despite the increase
in ME coefficient for AlScN based sensor, the LOD did not get
a similar boost at resonance. This was caused by the scaling
up thermomechanical noise with larger sensor deflection in
AlScN based sensor. It was also reported the working fre-
quency of such sensors is tunable by an external applied DC
electric field to the piezoelectric phase.
In 2013, Nan [131] reported a 215 MHz NEMS resonator to
detect DC magnetic field based on E effect. The structure of
the resonator is shown in Fig. 13(a). The resonator adopted an
AlN/(FeGaB/Al2O3)×10 ME heterostructure with interdigital
Pt electrode at bottom. The FeGaB soft magnetic multilayer
has lower eddy current loss compared with one single layer
and the interdigital electrode with designed pitch width drives
the resonator in extension vibration mode. With fixing pitch
width and equivalent density, the resonance frequency of the
sensor is purely determined by the effective Young’s modules,
which is affected by the DC magnetic field under detect due
to the E effect. Fig. 13(b) shows the admittance spectrum
under different applied DC magnetic field, presenting the
effect of changing Young’s modules electrically. The peak
admittance and quality factor Q under different applied DC
magnetic field is plotted in Fig. 13(c). The quality factor has
a minimum at 15 Oe, which is caused by the variation of
the magnetic domain wall state and the corresponding loss.
With large magnetic field applied, the magnetic domain wall is
eliminated, reducing the magnetic loss, and hence increasing
the quality factor. To test the sensing capability, the resonator
was biased under 5 Oe DC magnetic field where the reso-
nance frequency was at the most sensitive point to the applied
field. The resonator was then excited at a single frequency
for admittance measurement under a superimposed tiny DC
field as low as 50 pT. The measurement result is shown in
Fig. 13(d), where a 300 pT LOD was determined. The 10-
layer FeGaB/Al2O3magnetostrictive material also showed a
14 VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
FIGUR E 12. (a) Schematic of the MEMS AlScN resonant magnetoelectric sensor. (b) photograph of the packaged ME chip on a test board. (c) Measured
ME coefficient αME of AlScN and AlN based ME sensors at their mechanical resonance. (d) Voltage noise density spectra for AlScN (blue) and AlN
(orange) based ME sensors as well as a reference channel (black). (Reproduced from [11]).
self-bias performance which offered a LOD of 600 pT with no
DC magnetic field bias.
B. INTEGRATED PERMANENT MAGNET-BASED SENSOR
The permanent magnets represent a group of ferromagnetic
or ferrimagnetic materials that exhibit strong magnetization
after the removal of external field. The most commonly used
permanent hard magnet materials include ferrites, transition
metal alloys and rare-earth alloys [137]. Unlike the bulk per-
manent magnets fabrication, the silicon compatible integra-
tion of permanent magnets suffers the following difficulties
and challenges: 1). The harsh fabrication conditions including
the high temperature and pressure are not suitable for on-
chip process; 2). The traditional deposition methods includ-
ing electrodeposition [138], magnetron sputtering [139] and
pulsed laser deposition (PLD) [140] typically generate thin
film material with the thickness in the range of nanometer
to micrometer, which fairly limits the sample performance;
3). The wafer-level patterning of the as-deposited permanent
magnetic material may be hard to conduct or require specific
etching process that would damage the pre-exist structures.
Numerous efforts have been put into developing new powder-
based on-chip permanent magnet fabrication methodologies
to overcome these challenges including spin casting/screen
printing [141], dry-packing [142] and 3-dimensional (3D)-
printing [143]. The Ongoing research on powder-based per-
manent magnet integration is focusing on improving the pack-
ing density and the micro-structure of the magnetic powder
which is essential to the magnet performance.
With the advance of fabrication technique, the integrated
permanent magnet is utilized and serves as a sensing part
that responds to the external magnetic field under detect. The
mechanical force or torque generated is transferred to piezo-
electric phase and converted to electric outputs for readout
[116]–[118]. Compared to optical detection [119], [120], the
piezoelectric readout scheme does not need the optical gener-
ation, coupling and collimation instrument, which is favorable
for outdoor detection in harsh environment. In 2019, Niekiel
[118] reported a fully integrated permanent magnet-based sen-
sor that adopted the cantilever structure. The schematic of sen-
sor working principle and the structure are shown in Fig. 14(a)
and (b) respectively. When the integrated NdFeB magnet was
magnetized along the vertical direction and with the magnetic
field under detect applied along the horizontal direction, a
torque can be generated and drive the cantilever whose de-
flection was monitored by the AlN piezoelectric layer. Two
types of devices with different cantilever length are shown in
Fig. 14(c), where the electrodes were wire-bonded to a PCB
for the measurement. The sensitivity frequency spectrums are
VOLUME 1, NO. 4, OCTOBER 2021 15
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
FIGUR E 13. (a) Schematic of NEMS ME resonator. (b) Admittance curves of the NEMS sensor at various bias DC magnetic fields. (c) Resonance frequency
and admittance amplitude at the resonance frequency as a function of DC magnetic field. (d) Sensitivity measurement results under 5 Oe DC magnetic
field bias. (Reproduced from [131]).
plotted in Fig. 14(d). The peak sensitivity of sensor 1 and 2 are
34.6 kV/T and 37.1 kV/T, respectively. The sensors showed
a high quality factor oscillator characteristic which boosted
the sensitivity in band. The vertically magnetizing permanent
magnet configuration also offers superior sensitivity perfor-
mance compared to its horizontal counterparts which rely on
the magnetic field gradient for sensing. To determine the LOD
of the sensors, the electric noise density was measured and
divided by the sensitivity, it is shown in fig. 14(e) that the LOD
at resonance is 7.2 pT/Hz1/2 for sensor 1 and 7.3 pT/Hz1/2 for
sensor 2.
C. INTEGRATED MAGNETORESISTVE MATERIALS AND
SENSORS
Magnetoresistance effect refers to the effect that the resistance
of the thin film material varies under the external magnetic
field. The 3 major categories of MR including anisotropic
magnetoresistance (AMR), giant magnetoresistance (GMR)
and tunnel magnetoresistance (TMR). The AMR is attributed
to the anisotropic differences in the scattering of s electrons
off d electrons resulting from spin-orbit interaction [144] and
always happens in the transition metal and transition metal
alloys [145]. The GMR and TMR effect rely on the spin-
dependent electronic transport. All the MR effects have been
utilized for sensing applications and commercialized prod-
ucts are available. Typically, the MR sensors are used in
magnetic read head in magnetic hard disk drive, automotive
and consumer electronics including current leakage detection,
positioning, and biotechnology, etc [146]. The noise energy
spectrums of commercially available MR sensor products
are shown in Fig. 15 [147]. Note the term ’detectivity’ as
the Y-axis name in the plot refers to the noise floor level
. The sensor types corresponding to the product part numbers
are marked in the caption of the plot. It is shown among all the
MR sensors, AMR sensor has lowest noise performance in the
low frequency range. However, due to the physical nature of
these MR effect, the AMR have the lowest magnetoresistance
change, namely around 2%4% [144], compared to 10% in
GMR device under low magnetic field, and as high as 220%
[148], [149] to 600% [150] in TMR structures.
A typical GMR device consists of two ferromagnetic layers
with a non-magnetic conductive layer in between, whereas
for the TMR case, the non-magnetic layer in between is a
semiconductor or an isolator layer and the electron travels
through the layered structure by tunneling effect. The two
ferromagnetic layers are deposited to a designed thickness
such that a spontaneous anti-parallelly aligned magnetization
is achieved in these two layers. The external applied field
forces the magnetization in both ferromagnetic layers, which
subsequently decreases the overall resistance of the structure
due to the spin-dependent electron transport. The theory of the
spin-induced electric difference is elaborated by two current
16 VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
FIGUR E 14. (a) Working principle of the integrated permanent magnet-based sensor. (b) Cross-sectional view of the sensor structure. (c) Images of the
sensors mounted on PCB. (d) Frequency dependency of sensor sensitivity. (e) Frequency dependency of limit of detection. (Reproduced from [118]).
FIGUR E 15. The noise energy density of commercially available MR
sensors, in which NVE AAL002 is GMR sensor; NVE SDT is TMR sensor;
Honeywell HMC1001 is AMR sensor. (Reproduced from [141]).
models proposed by Mott [151]. Since the first discovery of
GMR effect by Fert in 1988 using Fe/Cr superlattice [152],
numerous efforts have been put into improving the magnetore-
sistance changing ratio of GMR/TMR thin film layered struc-
ture. In 1991, Dieny [153] proposed a structure named spin
valve where the magnetization of one of ferromagnetic layers
mentioned above was pined by an anti-ferromagnetic layer
(typically Mn alloy) adjacent to it through the spin interaction,
which overcame the issue that the previous structures were
only sensitive to high magnetic field. To further reduce the
field required to reverse the magnetization direction in ferro-
magnetic layers, a synthetic anti-ferromagnetic layer structure
that consists of two ferromagnetic layer (typically FeCo alloy)
and a coupling Ru layer was proposed by Parkin [154]. It was
found then by Parkin [155] introducing a Co-rich alloy layer
between the ferromagnetic layer and spacer of simple GMR
structure doubled the MR changing amount. The TMR effect
under room temperature was first demonstrated by Moodera
[156] in FeCo/Al2O3/Co structure where Al2O3acted as in-
sulting barrier for tunneling effect. The MgO isolation layer
was also widely used since then.
Although the GMR/TMR effect offers great potential in
sensing application thanks to their large MR change under
external magnetic field, it has some major drawbacks in-
cluding the non-linear response and hysteresis characteristic.
Typically, a DC magnetic bias generated by the Helmholtz
coils or permanent magnet can mitigate the issue [157]. Dif-
ferential [114] and bridge [115] sensor configurations have
VOLUME 1, NO. 4, OCTOBER 2021 17
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
FIGUR E 16. (a,b) Microscopy image of a serial TMR sensor. (c) Stacking structure of TMR sensor. (d) Outputs for one serial TMR sensor and (e) four serial
TMR sensors connected in a full Wheatstone bridge circuit at room temperature. (f) Noise spectrum of one serial TMR sensor and four serial TMR sensors
connected in a full Wheatstone bridge circuit. (Reproduced from [115]).
also been adopted to improve the sensitivity and diminish the
noise.
In [115], Jin fabricated and utilized the on-chip TMR sen-
sors for magnetic field leakage detection. The optical images
of the as-fabricated devices connected in series are shown
in Fig. 16(a) and (b), and the layered sensor structure with
constituent and thicknesses in nanometer of each layer is
schematically shown in Fig. 16(c). The TMR sensor adopted
synthetic ferromagnetic layers and the magnetization of the
upper one was pinned by IrMn material. The MgO layer was
adopted as tunneling barrier. The magnetic field sensitivity
was tested for one serial TMR sensor and four serial TMR
sensors connected in Wheatstone bridge configuration are
shown in Fig. 16(d) and (e), respectively. The sensitivity of
one sensor was 0.49 V/Oe and that of the 4-element bridge
is 0.37 V/Oe at zero field. Although some compromises were
made on the sensitivity, the 4-element bridge showed better
linear range by 6 Oe and nearly one order lower intensity in
noise energy spectrum which is shown in Fig. 16(f).
V. CONCLUSION
During the past several decades, integrated magnetic and mag-
netoelectric materials are widely utilized and enabling high-
efficiency, compact and tunable on-chip devices, which shows
a bright future for power, RF, microwave and sensing applica-
tions. Ongoing research is focusing on controlling the mag-
netic damping to either achieving high efficiency in magneto-
static wave based high frequency devices, or to control the fer-
romagnetic behavior and consequently domain wall dynamics
and spin transfer torque switching [158]. The integration of
highly dense hard magnet has long been another difficulty for
the community which is in the needs of more versatile fab-
rication methodologies with accurate positioning and defin-
able pattern on silicon. As for the integrated magnetoelectric,
surely the materials with high piezoelectric and magnetostric-
tive coefficients are highly demanded for stronger ME cou-
pling. The stress elimination in the two-phase ME composite
is also desirable to optimize the device performance, which
requires deposition condition, releasing process control and
even the post-fabrication annealing. Some researchers also
commit to get better understanding on physical basis of strain-
mediated magnetization and the radiation mechanism of ME
antennas. For practical applications, biomedical stimulus and
sensing incorporating the NEMS ME sensors and antennas
has become a research hotspot. Considering the miniaturiza-
tion brought by the novel ME devices on transmitting and
sensing devices, they are good candidates for invasive probing
and non-invasive wearable products in healthcare domain.
REFERENCES
[1] H. Szymczak, “Magnetic materials and applications,” in Encyclopedia
of Condensed Matter Physics. Waltham, MA, USA: Academic Press,
2005, pp. 204–211.
[2] A. Haldar, D. Kumar, and A. O. Adeyeye, “A reconfigurable wave-
guide for energy-efficient transmission and local manipulation of in-
formation in a nanomagnetic device,” Nature Nanotechnol., vol. 11,
no. 5, pp. 437–443, 2016.
[3] Y. Gao et al., “ Significantly enhanced inductance and quality factor
of GHz integrated magnetic solenoid inductors with FeGaB/Al2O3
18 VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
multilayer films,” IEEE Trans. Electron Devices, vol. 61, no. 5,
pp. 1470–1476, May 2014.
[4] Y. He et al., “Integrated tunable bandstop filter using self-biased
FeGaB/Al2O3multilayer thin film,” IEEE Trans. Magn., vol. 54, no. 9,
pp. 1–4, Sep. 2018.
[5] W. Che, E. K. N. Yung, K. Wu, and X. Nie, “Design investigation
on millimeter-wave ferrite phase shifter in substrate integrated wave-
guide,” Progr. Electromagn. Res., vol. 45, pp. 263–275, 2004.
[6] R. Zand, A. Roohi, and R. F. DeMara, “Energy-efficient and process-
variation-resilient write circuit schemes for spin hall effect MRAM
device,IEEE Trans. Very Large Scale Integr. (VLSI) Syst., vol. 25,
no. 9, pp. 2394–2401, Sep. 2017.
[7] D. E. Nikonov and G. I. Bourianoff, “Spin gain transistor in ferromag-
netic semiconductors-the semiconductor Bloch-equations approach,”
IEEE Trans. Nanotechnol., vol. 4, no. 2, pp. 206–214, Mar. 2005.
[8] S. Kaka, M. R. Pufall, W. H. Rippard, T. J. Silva, S. E. Russek, and
J. A. Katine, “Mutual phase-locking of microwave spin torque nano-
oscillators,” Nature, vol. 437, no. 7057, pp. 389–392, 2005.
[9] S. Gupta, P. Bhatter, and V. Kakkar, “Point-of-care detection of
tuberculosis using magnetoresistive biosensing chip,Tuberculosis,
vol. 127, 2021, Art. no. 102055.
[10] N. A. Spaldin and M. Fiebig, “The renaissance of magnetoelectric
multiferroics,” Science, vol. 309, no. 5733, pp. 391–392, 2005.
[11] J. Su et al., “AlScN-based MEMS magnetoelectric sensor,Appl.
Phys. Lett., vol. 117, no. 13, 2020, Art. no. 132903.
[12] S. K. Ghosh et al., “Rollable magnetoelectric energy harvester as a
wireless IoT sensor,” ACS Sustain. Chem. Eng., vol. 8, no. 2, 2019,
pp. 864–873.
[13] H. Chen et al., “Integrated tunable magnetoelectric RF inductors,”
IEEE Trans. Microw. Theory Techn., vol. 68, no. 3, 951–963,
Mar. 2020.
[14] H. Lin et al., “Integrated non-reciprocal dual H- and E-field tunable
bandpass filter with ultra-wideband isolation,” in Proc. IEEE MTT-S
Int. Microw. Symp., Phoenix, AZ, USA, May 2015, pp. 1–4.
[15] T. Nan et al., “Acoustically actuated ultra-compact NEMS magneto-
electric antennas,” Nature Commun., vol. 8, 2017, Art. no. 296.
[16] I. E. Dzyaloshinskii, “On the magneto-electrical effects in antiferro-
magnets,” Sov. Phys. JETP, vol. 10, pp. 628–629, 1960.
[17] N. A. Hill, “Why are there so few magnetic ferroelectrics?,J. Phys.
Chem. B, vol. 104, no. 29, pp. 6694–6709, 2000.
[18] X. Liang et al., “A review of thin-film magnetoelastic materials for
magnetoelectric applications,” Sensors, vol. 20, no. 5, 2020, Art.
no.1532.
[19] M. M. Vopson, Y. K. Fetisov, G. Caruntu, and G. Srinivasan, “Mea-
surement techniques of the magneto-electric coupling in multifer-
roics,” Materials, vol. 10, no. 8, 2017, Art. no. 963.
[20] S. Dong, J. Zhai, J. Li, and D. Viehland, “Near-ideal magneto-
electricity in high-permeability magnetostrictive/piezofiber laminates
with a (2-1) connectivity,Appl. Phys. Lett., vol. 89, no. 25, 2006,
Art. no. 252904.
[21] J. Lou, M. Liu, D. Reed, Y. Ren, and N. X. Sun, “Giant electric field
tuning of magnetism in novel multiferroic FeGaB/lead zinc niobate–
lead titanate (PZN-PT) heterostructures,” Adv. Mater., vol. 21, no. 46,
pp. 4711–4715, 2009.
[22] M. Liu et al., “Giant electric field tuning of magnetic properties in
multiferroic ferrite/ferroelectric heterostructures,” Adv. Funct. Mater.,
vol. 19, no. 11, pp. 1826–1831, 2009.
[23] M. Liu, S. Li, O. Obi, J. Lou, S. Rand, and N. X. Sun, “Electric field
modulation of magnetoresistance in multiferroic heterostructures for
ultralow power electronics,Appl. Phys. Lett., vol. 98, no. 22, 2011,
Art. no. 222509.
[24] J. M. Hu, Z. Li, L. Q. Chen, and C. W. Nan, “Design of a voltage-
controlled magnetic random access memory based on anisotropic
magnetoresistance in a single magnetic layer,” Adv. Mater., vol. 24,
no. 21, pp. 2869–2873, 2012.
[25] J. H. Griffiths, “Anomalous high-frequency resistance of ferromag-
netic metals,” Nature, vol. 158, no. 4019, pp. 670–671, 1946.
[26] C. Kittel, “Ferromagnetic resonance,” J. Phys. Radium, vol. 12, no. 3,
pp. 291–302, 1951.
[27] N. X. Sun and G. Srinivasan, Voltage Control of Magnetism in Mul-
tiferroic Heterostructures and Devices. Singapore: World Scientific,
vol. 2, no. 3, 2012, Art. no. 1240004.
[28] D. Berlincourt, H. H. A. Krueger, and C. Near “Technical publication
TP-226, properties of piezoelectricity ceramics,” Morgan Electro Ce-
ramics, vol. 12, 2003.
[29] J. Erhart, P. P˚
ulpán, and M. Pustka, Piezoelectric Ceramic Resonators.
Cham, Switzerland: Springer, 2017.
[30] C. Tu et al., “Mechanical-resonance-enhanced thin-film magnetoelec-
tric heterostructures for magnetometers, mechanical antennas, tun-
able RF inductors, and filters,” Materials, vol. 12, no. 14, 2019,
Art. no. 2259.
[31] K. Tsubouchi; K. Sugai, and N. Mikoshiba, “AlN material constants
evaluation and SAW properties on AlN/Al2O3and AlN/Si,” in Proc.
Ultrason. Symp., Chicago, IL, USA, Oct. 1981, pp. 375–380.
[32] M.-A. Dubois, and P. Muralt, “Properties of aluminum nitride thin
films for piezoelectric transducers and microwave filter applications,
Appl. Phys. Lett., vol. 74, pp. 3032–3034, 1999.
[33] J. - L. Sanchez-Rojas et al., “Advanced determination of piezoelectric
properties of AlN thin films on silicon substrates,” in Proc. IEEE
Ultrason. Symp., Beijing, China, Nov. 2008, pp. 903–906.
[34] L. C. Lim, K. K. Rajan, and J. Jin, “Characterization of flux-grown
PZN-PT single crystals for high-performance piezo devices,” IEEE
Trans. Ultrason. Ferroelectr. Freq. Control, vol. 54, no. 12, pp. 2474–
2478, Dec. 2007.
[35] P. Han, W. Yan, J. Tian, X. Huang, and H. Pan, “Cut directions for the
optimization of piezoelectric coefficients of lead magnesium niobate–
lead titanate ferroelectric crystals,” Appl. Phys. Lett., vol. 86, no. 5,
2005, Art. no. 052902.
[36] K. Uchino, Advanced Piezoelectric Materials: Science and Technol-
ogy. Sawston, U.K.: Woodhead, 2017.
[37] Z. Yang and J. Zu, “Comparison of PZN-PT, PMN-PT single crystals
and PZT ceramic for vibration energy harvesting,Energy Convers.
Manage., vol. 122, pp. 321–329, 2016.
[38] G. Engdahl and I. D. Mayergoyz, Handbook of Giant Magnetostrictive
Materials, New York, NY, USA: Academic, 2000.
[39] A. E. Clark, M. Wun-Fogle, J. B. Restorff, and T. A. Lograsso, “Mag-
netostrictive properties of Galfenol alloys under compressive stress,
Mater. Trans., vol. 43, no. 5, pp. 881–886, 2002.
[40] N. Srisukhumbowornchai and S. Guruswamy, “Large magnetostriction
in directionally solidified FeGa and FeGaAl alloys,” J. Appl. Phys.,
vol. 90, no. 11, pp. 5680–5688, Nov. 2001.
[41] J. Lou, R. E. Insignares, Z. Cai, K. S. Ziemer, M. Liu and N. X.
Sun, “Soft magnetism, magnetostriction, and microwave properties of
FeGaB thin films,” Appl. Phys. Lett., vol. 91, no. 18, pp. 182–504, Oct.
2007.
[42] H. Greve, E. Woltermann, H.-J. Quenzer, B. Wagner, and E. Quandt,
“Giant magnetoelectric coeffcients in (Fe90Co10)78 Si12 B10-AlN
thin film composites,” Appl. Phys. Lett., vol. 96, 2010, Art. no. 182501.
[43] Z. Chu et al., “Enhanced resonance magnetoelectric coupling in (1-1)
connectivity composites,Adv. Mater., vol. 29, 2017, Art. no. 1606022.
[44] M. Liu et al., “Electrically induced enormous magnetic anisotropy
in Terfenol-D/lead zinc niobate-lead titanate multiferroic heterostruc-
tures,” J. Appl. Phys., vol. 112, no. 6, 2012, Art. no. 063917.
[45] A. Butera, J. Gómez, J. L. Weston, and J. A. Barnard, “Growth and
magnetic characterization of epitaxial Fe 81 Ga 19/MgO (100) thin
films,” J. Appl. Phys., vol. 98, no. 3, 2005, Art. no. 033901.
[46] C. L. Platt, N. K. Minor, and T. J. Klemmer, “Magnetic and structural
properties of FeCoB thin films,” IEEE Trans. Magn., vol. 37, no. 4,
pp. 2302–2304, Apr. 2001.
[47] I. Kim, J. Kim, K. H. Kim, and M. Yamaguchi, “Effects of boron
contents on magnetic properties of Fe-Co-B thin films,” IEEE Trans.
Magn., vol. 40, no. 4, pp. 2706–2708, Apr. 2004.
[48] J. Lou, R. E. Insignares, Z. Cai, K. S. Ziemer, M. Liu, and N. X.
Sun, “Soft magnetism, magnetostriction, and microwave properties of
FeGaB thin films,” Appl. Phys. Lett., vol. 91, no. 18, 2007, p.182504.
[49] C. Dong et al., “Characterization of magnetomechanical properties
in FeGaB thin films,” Appl. Phys. Lett., vol. 113, no. 26, 2018,
Art. no. 262401.
[50] Z. Chu, M. PourhosseiniAsl, and S. Dong, “Review of multi-layered
magnetoelectric composite materials and devices applications,” J.
Phys.D,Appl.Phys.
, vol. 51, no. 24, 2018, Art. no. 243001.
[51] W. Ling, C. Xiaoning, X. Hang, H. Wenmei, W. Bowen, and Y. Zeyu,
“Frequency-dependent complex magnetic permeability and magnetic
losses of Fe-Ga alloy,” in Proc. IEEE 20th Int. Conf. Elect. Machines
Syst., Aug. 2017, pp. 1–5.
VOLUME 1, NO. 4, OCTOBER 2021 19
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
[52] Z. Zhang et al., “Coupling of magneto-strictive FeGa film with single-
crystal diamond MEMS resonator for high-reliability magnetic sens-
ing at high temperatures,” Mater. Res. Lett., vol. 8, no. 5, pp. 180–186,
2020.
[53] X. Wang et al., “Size-dependent magnetic properties of FeGaB/Al2O3
multilayer micro-islands,” Phys. Lett. A, vol. 382, no. 23, pp. 1505–
1508, 2018.
[54] F.Q.Xiao,M.G.Han,H.P.Lu,L.Zhang,E.Li,andL.J.Deng,
“Out-of-plane magnetic anisotropy and microwave permeability of
magnetoelastic FeCoSiB amorphous thin films,” J. Electron. Sci. Tech-
nol., vol. 5, no. 3, pp. 226–229, 2007.
[55] E. J. Yun, M. Jung, C. I. Cheon, and H. G. Nam, “Microfabrication
and characteristics of low-power high-performance magnetic thin-film
transformers,” IEEE Trans. Magn., vol. 40, no. 1, pp. 65–70, Jan. 2004.
[56] B. M. Mundotiya, D. Dinulovic, L. Rissing, and M. C. Wurz, “Fab-
rication and characterization of a Ni-Fe-W core microtransformer
for high-Frequency power applications,Sens. Actuators A, Phys.,
vol. 267, pp. 42–47, 2017.
[57] S. G. Mariappan, A. Moazenzadeh, and U. Wallrabe, “Polymer mag-
netic composite core based microcoils and microtransformers for very
high frequency power applications,Micromachines, vol. 7, no. 4,
2016, Art. no. 60.
[58] X. Wang et al., “A novel NiZn ferrite integrated magnetic solenoid
inductor with a high quality factor at 0.7–6 GHz,” AIP Adv.,vol.7,
no. 5, 2017, Art. no. 056606.
[59] H. Kurata, K. Shirakawa, O. Nakazima, and K. Murakami, “Study of
thin film micro transformer with high operating frequency and cou-
pling coefficient,IEEE Trans. Magn., vol. 29, no. 6, pp. 3204–3206,
Nov. 1993.
[60] D. Flynn, A. Toon, L. Allen, R. Dhariwal, and M. P. Desmulliez,
“Characterization of core materials for microscale magnetic com-
ponents operating in the megahertz frequency range,IEEE Trans.
Magn., vol. 43, no. 7, pp. 3171–3180, Aug. 2007.
[61] Y. Gao et al., “Power-efficient voltage tunable RF integrated magneto-
electric inductors with FeGaB/Al2O3multilayer films,” in Proc. IEEE
MTT-S Int. Microw. Symp., 2014, pp. 1–4.
[62] J. Mullenix, A. El-Ghazaly, and S. X. Wang, “Integrated transformers
with sputtered laminated magnetic core,” IEEE Trans. Magn., vol. 49,
no. 7, pp. 4021–4027, Jul. 2013.
[63] A. El-Ghazaly, R. M. White, and S. X. Wang, “Gigahertz-band in-
tegrated magnetic inductors,” IEEE Trans. Microw. Theory Techn.,
vol. 65, no. 12, pp. 4893–4900, Dec. 2017.
[64] C. Kittel, “On the theory of ferromagnetic resonance absorption,”
Phys. Rev., vol. 73, no. 2, 1948, Art. no. 155.
[65] M. Yamaguchi, S. Bae, K. H. Kim, K. Tan, T. Kusumi, and K. Ya-
makawa, “Ferromagnetic RF integrated inductor with closed magnetic
circuit structure,” in IEEE MTT-S Int. Microw. Symp. Dig., Jun. 2005,
pp. 351–354.
[66] M. Yamaguchi et al., “Magnetic RF integrated thin-film inductors,” in
IEEE MTT-S Int. Microw. Symp. Dig., Jun. 2000, pp. 205–208.
[67] J. Salvia, J. A. Bain, and C. P. Yue, “Tunable on-chip inductors up
to 5 GHz using patterned permalloy laminations,” in Proc. IEEE Int.
Electron Devices Meeting, Dec. 2005, pp. 943–946.
[68] D. S. Gardner, G. Schrom, F. Paillet, B. Jamieson, T. Karnik, and S.
Borkar, “Review of on-chip inductor structures with magnetic films,
IEEE Trans. Magn., vol. 45, no. 10, pp. 4760–4766, Oct. 2009.
[69] P. Park, C. S. Kim, M. Y. Park, S. D. Kim, and H. K. Yu, “Variable
inductance multilayer inductor with MOSFET switch control,” IEEE
Electron Device Lett., vol. 25, no. 3, pp. 144–146, Mar. 2004.
[70] A. Shirane, H. Ito, N. Ishihara, and K. Masu, “Planar solenoidal induc-
tor in radio frequency micro-electro-mechanical systems technology
for variable inductor with wide tunable range and high quality factor,
Jpn. J. Appl. Phys., vol. 51, no. 5S, 2012, Art. no. 05EE02.
[71] J. I. Kim and D. Peroulis, “Tunable MEMS spiral inductors with opti-
mized RF performance and integrated large-displacement electrother-
mal actuators,” IEEE Trans. Microw. Theory Techn., vol. 57, no. 9,
pp. 2276–2283, Oct. 2009.
[72] I. Zine-El-Abidine, M. Okoniewski, and J. G. McRory, “RF MEMs
tunable inductor using bimorph microactuators,” in Proc. IEEE Int.
Conf. MEMS, NANO Smart Syst., Jul. 2005, pp. 436–437.
[73] A. Bhattacharya, D. Mandal, and T. K. Bhattacharyya, “A 1.3–2.4-
GHz 3.1-mW VCO using electro-thermo-mechanically tunable self-
assembled MEMS inductor on HR substrate,” IEEE Trans. Microw.
Theory Techn., vol. 63, no. 2, pp. 459–469, Feb. 2015.
[74] D. M. Fang, Q. Yuan, X. H. Li, and H. X. Zhang, “Electrostatically
driven tunable radio frequency inductor,Microsyst. Technol., vol. 16,
no. 12, 2010, pp. 2119–2122.
[75] M. Vroubel, Y. Zhuang, B. Rejaei, and J. N. Burghartz, “Integrated
tunable magnetic RF inductor,” IEEE Electron Device Lett., vol. 25,
no. 12, pp. 787–789, Jan. 2005.
[76] T. Wang et al., “Novel electrically tunable microwave solenoid induc-
tor and compact phase shifter utilizing permaloy and PZT thin films,”
IEEE Trans. Microw. Theory Techn., vol. 65, no. 10, pp. 3569–3577,
Oct. 2017.
[77] T. Wang et al., “Integrating nanopatterned ferromagnetic and fer-
roelectric thin films for electrically tunable RF applications,” IEEE
Trans. Microw. Theory Techn., vol. 65, no. 2, pp. 504–512, Feb. 2016.
[78] J. Lou, D. Reed, M. Liu, and N. X. Sun, “Electrostatically tun-
able magnetoelectric inductors with large inductance tunability,Appl.
Phys. Lett., vol. 94, no. 11, 2009, Art. no. 112508.
[79] H. Lin et al., “Voltage tunable magnetoelectric inductors with im-
proved operational frequency and quality factor for power electronics,
IEEE Trans. Magn., vol. 51, no. 1, pp. 1–5, Jan. 2014.
[80] S. Jiang, Y. Liu, Z. Mei, J. Peng, and C. M. Lai, “A magnetic integrated
LCL–EMI filter for a single-phase SiC-MOSFET grid-connected in-
verter,IEEE J. Emerg. Sel. Topics Power Electron., vol. 8, no. 1,
pp. 601–617, Mar. 2020.
[81] J. Fang, H. Li, and Y. Tang, “A magnetic integrated LLCL filter for
grid-connected voltage-source converters,IEEE Trans. Power Elec-
tron., vol. 32, no. 3, pp. 1725–1730, Sep. 2016.
[82] W. Wu, C. S. Tsai, C. C. Lee, H. J. Yoo, H. Hopster, and D. L.
Mills, “Ferromagnetic Fe/Ag-gaas waveguide structures for wideband
microwave integrated notch filter devices,” in 58th DRC, Device Res.
Conf., Dig., Jun. 2000, pp. 51–52.
[83] I. Harward, R. E. Camley, and Z. Celinski, “On-wafer magnetically
tunable millimeter wave notch filter using M-phase Ba hexagonal
ferrite/Pt thin films on Si,” Appl. Phys. Lett., vol. 105, no. 17, 2014,
Art. no.173503.
[84] C. S. Tsai et al., “Tunable wideband microwave band-stop and band-
pass filters using YIG/GGG-GaAs layer structures,” IEEE Trans.
Magn., vol. 41, no. 10, pp. 3568–3570, Nov. 2005.
[85] Y. Y. Song, C. L. Ordóñez-Romero, and M. Wu, “Millimeter wave
notch filters based on ferromagnetic resonance in hexagonal barium
ferrites,” Appl. Phys. Lett., vol. 95, no. 14, 2009, Art. no. 142506.
[86] H. Lin et al., “Tunable RF band-pass filters based on NEMS mag-
netoelectric resonators,” in Proc. IEEE MTT-S Int. Microw. Symp.,
May 2016, pp. 1–4.
[87] K. Entesari and G. M. Rebeiz, “A differential 4-bit 6.5-10-GHz RF
MEMS tunable filter,” IEEE Trans. Microw. Theory Techn., vol. 53,
no. 3, pp. 1103–1110, Mar. 2005.
[88] C. C. Cheng and G. M. Rebeiz, “High Q 4–6-GHz suspended stripline
RF MEMS tunable filter with bandwidth control,” IEEE Trans. Mi-
crow. Theory Techn., vol. 59, no. 10, pp. 2469–2476, Oct. 2011.
[89] V. Sekar, M. Armendariz, and K. Entesari, “A 1.2–1.6-GHz substrate-
integrated-waveguide RF MEMS tunable filter,IEEE Trans. Microw.
Theory Techn., vol. 59, no. 4, pp. 866–876, Apr. 2011.
[90] S. M. Hanna and S. Zeroug, “Single and coupled MSW resonators
for microwave channelizers,IEEE Trans. Magn., vol. TSM-24, no. 6,
pp. 2808–2810, Jun. 1988.
[91] W. S. Ishak and K. W. Chang, “Tunable microwave resonators us-
ing magnetostatic wave in YIG films,IEEE Trans. Microw. Theory
Techn., vol. MTT-34, no. 12, pp. 1383–1393, Dec. 1986.
[92] J. Wu, X. Yang, S. Beguhn, J. Lou, and N. X. Sun, “Nonreciprocal
tunable low-loss bandpass filters with ultra-wideband isolation based
on magnetostatic surface wave,IEEE Trans. Microw. Theory Techn.,
vol. 60, no. 12, pp. 3959–3968, Dec. 2012.
[93] D. B. Miron, Small Antenna Design. Amsterdam, The Netherlands;
New York, NY, USA: Elsevier, 2006.
[94] J. Volakis, C. C. Chen, and K. Fujimoto, Small Antennas: Miniaturiza-
tion Techniques & Applications. New York, NY, USA: McGraw Hill,
2009.
[95] O. Staub, J. F. Zurcher, A. K. Skrivervik, and J. R. Mosig, “PCS
antenna design: The challenge of miniaturisation,” in Proc. IEEE An-
tennas Propag. Soc. Int. Symp./USNC/URSI Nat. Radio Sci. Meeting,
Jul. 1999, pp. 548–551.
[96] C. Pfeiffer, “Fundamental efficiency limits for small metallic anten-
nas,” IEEE Trans. Antennas Propag., vol. 65, no. 4, pp. 1642–1650,
Apr. 2017.
20 VOLUME 1, NO. 4, OCTOBER 2021
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
[97] L. J. Chu, “Physical limitations of omni-directional antennas,” J. Appl.
Phys., vol. 19, no. 12, pp. 1163–1175, 1948.
[98] R. F. Harrington, “Effect of antenna size on gain, bandwidth, and
efficiency,” J. Res. Nat. Bur. Standards D, Radio Propag., vol. 64,
no. 1, pp. 1–12, 1960.
[99] R. C. Hansen, “Fundamental limitations in antennas,” Proc. IEEE,
vol. 69, no. 2, pp. 170–182, 1981.
[100] Z. Yao, Y. E. Wang, S. Keller, and G. P. Carman, “Bulk acoustic
wave-mediated multiferroic antennas: Architecture and performance
bound,” IEEE Trans. Antennas Propag., vol. 63, no. 8, pp. 3335–3344,
Aug. 2015.
[101] J. P. Domann and G. P. Carman, “Strain powered antennas,” J. Appl.
Phys., vol. 121, no. 4, 2017, Art. no. 044905.
[102] C. Dong et al., “A portable very low frequency (VLF) communica-
tion system based on acoustically actuated magnetoelectric antennas,”
IEEE Antennas Wireless Propag. Lett., vol. 19, no. 3, pp. 398–402,
Jan. 2020.
[103] T. Karacolak, A. Z. Hood, and E. Topsakal, “Design of a dual-band
implantable antenna and development of skin mimicking gels for
continuous glucose monitoring,” IEEE Trans. Microw. Theory Techn.,
vol. 56, no. 4, pp. 1001–1008, May 2008.
[104] T. Rupp, B. D. Truong, S. Williams, and S. Roundy, “Magnetoelectric
transducer designs for use as wireless power receivers in wearable and
implantable applications,” Materials, vol. 12, no. 3, 2019, Art. no. 512.
[105] F. J. Huang, C. M. Lee, C. L. Chang, L. K. Chen, T. C. Yo, and C.
H. Luo, “Rectenna application of miniaturized implantable antenna
design for triple-band biotelemetry communication,” IEEE Trans. An-
tennas Propag., vol. 59, no. 7, pp. 2646–2653, Jul. 2011.
[106] C. M. Lee, T. C. Yo, and C. H. Luo, “Compact broadband stacked
implantable antenna for biotelemetry with medical devices,” in
Proc. IEEE Annu. Wireless Microw. Technol. Conf., Dec. 2006,
pp. 1–4.
[107] M. Zaeimbashi et al., “NanoNeuroRFID: A wireless implantable de-
vice based on magnetoelectric antennas,” IEEE J. Electromagn., RF
Microw. Med. Biol., vol. 3, no. 3, pp. 206–215, Mar. 2019.
[108] H. Lin et al., “NEMS magnetoelectric antennas for biomedical ap-
plication,” in Proc. IEEE Int. Microw. Biomed. Conf., Jun. 2018,
pp. 13–15.
[109] D. Robbes, “Highly sensitive magnetometers—A review,Sens. Actu-
ators A, Phys., vol. 129, no. 1/2, pp. 86–93, 2006.
[110] M. Schmelz, et al., “Sub-fT/Hz1/2 resolution and field-stable SQUID
magnetometer based on low parasitic capacitance sub-micrometer
cross-type Josephson tunnel junctions,” Physica C, Supercond. Appl.,
vol. 482, pp. 27–32, 2012.
[111] B. Gojdka et al., “Fully integrable magnetic field sensor based
on delta-E effect,Appl. Phys. Lett., vol. 99, no. 22, 2011,
Art. no. 223502.
[112] Y. Hui, T. Nan, N. X. Sun, and M. Rinaldi, “High resolution mag-
netometer based on a high frequency magnetoelectric MEMS-CMOS
oscillator,J. Microelectromech. Syst., vol. 24, no. 1, pp. 134–143,
2014.
[113] M. Li et al., “Ultra-sensitive NEMS magnetoelectric sensor for pi-
cotesla DC magnetic field detection,” Appl. Phys. Lett., vol. 110,
no. 14, 2017, Art. no. 143510.
[114] C. Mu¸suroi, M. Oproiu, M. Volmer, and I. Firastrau, “High sensi-
tivity differential giant magnetoresistance (GMR) based sensor for
non-contacting DC/AC current measurement,Sensors, vol. 20, no. 1,
2020, Art. no. 323.
[115] Z. Jin, M. N. Sam, M. Oogane, and Y. Ando, “Serial MTJ-based
TMR sensors in bridge configuration for detection of fractured steel
bar in magnetic flux leakage testing,” Sensors, vol. 21, no. 2, 2021,
Art. no. 668.
[116] E. S. Leland, P. K. Wright, and R. M. White, “A MEMS AC current
sensor for residential and commercial electricity end-use monitoring,”
J. Micromech. Microeng., vol. 19, no. 9, 2009, Art. no. 094018.
[117] S. B. Lao, S. S. Chauhan, T. E. Pollock, T. Schröder, I. S. Cho, and
A. Salehian, “Design, fabrication and temperature sensitivity testing
of a miniature piezoelectric-based sensor for current measurements,”
Actuators, vol. 3, no. 3, pp. 162–181, 2014.
[118] F. Niekiel et al., “Highly sensitive MEMS magnetic field sensors
with integrated powder-based permanent magnets,Sens. Actuators A,
Phys., vol. 297, 2019, Art. no. 111560.
[119] L. Long and S. Zhong, “A MEMS torsion magnetic sensor with re-
flective blazed grating integration,J. Micromech. Microeng., vol. 26,
no. 7, 2016, Art. no. 075004.
[120] L. Long, M. Wang, and S. Zhong, “A torsion MEMS magnetic sen-
sor with permanent magnet and fiber-optic detection,” IEEE Sens. J.,
vol. 16, no. 23, pp. 8426–8433, Dec. 2016.
[121] S. O. Choi, S. Kawahito, Y. Matsumoto, M. Ishida, and Y. Tadokoro,
“An integrated micro fluxgate magnetic sensor,” Sens. Actuators A,
Phys., vol. 55, no. 2/3, pp. 121–126, 1996.
[122] Y. Wang, J. Li, and D. Viehland, “Magnetoelectrics for magnetic sen-
sor applications: Status, challenges and perspectives,Mater. Today,
vol. 17, no. 6, pp. 269–275, 2014.
[123] J. Gao, J. Das, Z. Xing, J. Li, and D. Viehland, “Comparison of noise
floor and sensitivity for different magnetoelectric laminates,J. Appl.
Phys., vol. 108, no. 8, 2010, Art. no. 084509.
[124] L. Shen et al., “Magnetoelectric nonlinearity in magnetoelectric lami-
nate sensors,” J. Appl. Phys., vol. 110, no. 11, 2011, Art. no. 114510.
[125] J. Das, J. Gao, Z. Xing, J. F. Li, and D. Viehland, “Enhancement in the
field sensitivity of magnetoelectric laminate heterostructures,Appl.
Phys. Lett., vol. 95, no. 9, 2009, Art. no. 092501.
[126] M. Li, D. Berry, J. Das, D. Gray, J. Li, and D. Viehland, “Enhanced
sensitivity and reduced noise floor in magnetoelectric laminate sensors
by an improved lamination process,J. Amer. Ceram. Soc., vol. 94,
no. 11, pp. 3738–3741, 2011.
[127] J. Zhai, Z. Xing, S. Dong, J. Li, and D. Viehland, “Detection of
pico-Tesla magnetic fields using magneto-electric sensors at room
temperature,” Appl. Phys. Lett., vol. 88, no. 6, 2006, Art. no. 062510.
[128] J. Gao, Y. Wang, M. Li, Y. Shen, J. Li, and D. Viehland, “Quasi-static
(f<102 Hz) frequency response of magnetoelectric composites
based magnetic sensor,Mater. Lett., vol. 85, pp. 84–87, 2012.
[129] P. Zhao et al., “Fabrication and characterization of all-thin-film
magnetoelectric sensors,” Appl. Phys. Lett., vol. 94, no. 24, 2009,
Art. no. 243507.
[130] E. Yarar et al., “Inverse bilayer magnetoelectric thin film sensor,Appl.
Phys. Lett., vol. 109, no. 2, 2016, Art. no. 022901.
[131] T. Nan, Y. Hui, M. Rinaldi, and N. X. Sun, “Self-biased 215MHz
magnetoelectric NEMS resonator for ultra-sensitive DC magnetic field
detection,” Sci. Rep., vol. 3, no. 1, pp. 1–6, 2013.
[132] R. Jahns, H. Greve, E. Woltermann, E. Quandt, and R. H. Knochel,
“Noise performance of magnetometers with resonant thin-film mag-
netoelectric sensors,” IEEE Trans. Instrum. Meas., vol. 60, no. 8,
pp. 2995–3001, Sep. 2011.
[133] A. A. Chlaihawi et al., “Novel screen printed flexible magnetoelectric
thin film sensor,” Procedia Eng., vol. 168, pp. 684–687, 2016.
[134] R. Jahns, H. Greve, E. Woltermann, E. Quandt, and R. Knöchel, “Sen-
sitivity enhancement of magnetoelectric sensors through frequency-
conversion,Sens. Actuators A, Phys., vol. 183, pp. 16–21, 2012.
[135] V. Röbisch et al., “Exchange biased magnetoelectric composites for
magnetic field sensor application by frequency conversion,J. Appl.
Phys., vol. 117, no. 17, 2015, Art. no. 17B513.
[136] S. Salzer et al., “Generalized magnetic frequency conversion for thin-
film laminate magnetoelectric sensors,” IEEE Sens. J., vol. 17, no. 5,
pp. 1373–1383, May 2016.
[137] D. P. Arnold and N. Wang, “Permanent magnets for MEMS,” J. Mi-
croelectromech. Syst., vol. 18, no. 6, pp. 1255–1266, 2009.
[138] O. Berkh, Y. Rosenberg, Y. Shacham-Diamand, and E. Gileadi, “De-
position of CoPtP films from citric electrolyte,” Microelectron. Eng.,
vol. 84, no. 11, pp. 2444–2449, 2007.
[139] A. Walther, D. Givord, N. M. Dempsey, K. Khlopkov, and O. Gut-
fleisch, “Structural, magnetic, and mechanical properties of 5 μm thick
smco films suitable for use in microelectromechanical systems,” J.
Appl. Phys., vol. 103, no. 4, 2008, Art. no. 043911.
[140] F. J. Cadieu, L. Chen, and B. Li, “Enhanced magnetic properties of
nanophase SmCo5 film dispersions,” IEEE Trans. Magn., vol. 37,
no. 4, pp. 2570–2572, Apr. 2001.
[141] S. Schwarzer, B. Pawlowski, A. Rahmig, and J. Töpfer, “Perma-
nent magnetic thick films from remanence optimized NdFeB-inks,” J.
Mater. Sci., Mater. Electron., vol. 15, no. 3, pp. 165–168, 2004.
[142] Y. He et al., “High-performance on-chip hot-pressed NdFeB hard
magnets for MEMS applications,” IEEE Trans. Magn., vol. 57, no. 4,
pp. 1–4, Feb. 2021.
[143] B. G. Compton et al., “Direct-write 3D printing of NdFeB bonded
magnets,” Mater. Manuf. Processes, vol. 33, no. 1, pp. 109–113, 2018.
VOLUME 1, NO. 4, OCTOBER 2021 21
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
HE ET AL.: INTEGRATED MAGNETICS AND MAGNETOELECTRICS FOR SENSING, POWER, RF, AND MICROWAVE ELECTRONICS
[144] L. Jiang, A. Gokce, F. C. S. Silva, and E. R. Nowak, “Zigzag-
shaped AMR magnetic sensors: Transfer characteristics and noise,
in Noise and Information in Nanoelectronics, Sensors, and Standards
III, Bellingham WA, USA: SPIE, May 2005, vol. 5846, pp. 156–168.
[145] H. Eberta, A. Vernesa, and J. Banhartb, “Magnetoresistance,
anisotropic,” Encyclopedia of Materials: Science and Technology. 2nd
ed., Oxford, U.K.: Pergamon Press, 2001, pp. 5079–5083.
[146] L. Jogschies et al., “Recent developments of magnetoresistive sensors
for industrial applications,” Sensors, vol. 15, no. 11, pp. 28665–28689,
2015.
[147] W. F. Egelhoff, et al., “Critical challenges for picoTesla magnetic-
tunnel-junction sensors,” Sens. Actuators A, Phys., vol. 155, no. 2,
pp. 217–225, 2009.
[148] S. Yuasa, T. Nagahama, A. Fukushima, Y. Suzuki, and K. Ando, “Gi-
ant room-temperature magnetoresistance in single-crystal Fe/MgO/Fe
magnetic tunnel junctions,” Nature Mater., vol. 3, no. 12, pp. 868–871,
2004.
[149] S. S. Parkin et al., “Giant tunnelling magnetoresistance at room tem-
perature with MgO (100) tunnel barriers,” Nature Mater., vol. 3, no. 12,
pp. 862–867, 2004.
[150] S. Ikeda et al., “Tunnel magnetoresistance of 604% at 300 K by sup-
pression of Ta diffusion in CoFeB/MgO/CoFeB pseudo-spin-valves
annealed at high temperature,” Appl. Phys. Lett., vol. 93, no. 8, 2008,
Art. no. 082508.
[151] N. F. Mott, “The resistance and thermoelectric properties of the tran-
sition metals,” Proc. Roy. Soc. Lond. A, Math. Phys. Sci., vol. 156,
no. 888, pp. 368–382, 1936.
[152] M. N. Baibich et al., “Giant magnetoresistance of (001) Fe/(001)
Cr magnetic superlattices,” Phys. Rev. Lett., vol. 61, no. 21, 1988,
Art. no. 2472.
[153] B. Dieny et al., “Magnetotransport properties of magnetically soft
spin-valve structures,J. Appl. Phys., vol. 69, no. 8, pp. 4774–4779,
1991.
[154] S. S. P. Parkin, N. More, and K. P. Roche, “Oscillations in exchange
coupling and magnetoresistance in metallic superlattice structures:
Co/Ru, Co/Cr, and Fe/Cr,Phys. Rev. Lett., vol. 64, no. 19, 1990,
Art. no. 2304.
[155] S. S. P. Parkin, “Origin of enhanced magnetoresistance of mag-
netic multilayers: Spin-dependent scattering from magnetic interface
states,” Phys. Rev. Lett., vol. 71, no. 10, 1993, Art. no. 1641.
[156] J. S. Moodera, L. R. Kinder, T. M. Wong, and R. Meservey, “Large
magnetoresistance at room temperature in ferromagnetic thin film tun-
nel junctions,” Phys. Rev. Lett., vol. 74, no. 16, 1995, Art. no. 3273.
[157] C. Mu¸suroi, M. Oproiu, M. Volmer, J. Neamtu, M. Avram, and E.
Helerea, “Low field optimization of a non-contacting high-sensitivity
GMR-based DC/AC current sensor,Sensors, vol. 21, no. 7, 2021,
Art. no. 2564.
[158] A. Barman, S. Mondal, S. Sahoo, and A. De, “Magnetization dynam-
ics of nanoscale magnetic materials: A perspective,J. Appl. Phys.,
vol. 128, no. 17, 2020, Art. no. 170901.
[159] M. Akiyama, T. Kamohara, K. Kano, A. Teshigahara, Y. Takeuchi, and
N. Kawahara, “Enhancement of piezoelectric response in scandium
aluminum nitride alloy thin films prepared by dual reactive cosputter-
ing,” Adv. Mater., vol. 21, no. 5, pp. 593–596, 2009.
[160] S. Fichtner, N. Wolff, F. Lofink, L. Kienle, and B. Wagner, “AlScN:
A III-V semiconductor based ferroelectric,” J. Appl. Phys., vol. 125,
no. 11, 2019, Art. no. 114103.
[161] F. Parsapour, V. Pashchenko, P. Nicolay, and P. Murali, “Material
constants extraction for AlScN thin films using a dual mode baw res-
onator,” in Proc. IEEE Micro Electro Mech. Syst., 2018, pp. 763–766.
[162] F. Parsapour et al., “Ex-situ AlN seed layer for (0001)-textured al 0.84
sc 0.16 n thin films grown on SiO2substrates,” in Proc. IEEE Int.
Ultrason. Symp., 2017, pp. 1–4.
[163] R. H. Olsson, Z. Tang, and M. D’Agati, “Doping of aluminum nitride
and the impact on thin film piezoelectric and ferroelectric device per-
formance,” in Proc. IEEE Custom Integr. Circuits Conf., Mar. 2020,
pp. 1–6.
[164] P. M. Mayrhofer, P. ˚
A. Persson, A. Bittner, and U. Schmid, “Properties
of Sc x Al 1-x N (x=0.27) thin films on sapphire and silicon substrates
upon high temperature loading,” Microsyst. Technol., vol. 22, no. 7,
pp. 1679–1689, 2016.
[165] S. Chauhan et al., “Multiferroic, magnetoelectric and optical prop-
erties of Mn doped BiFeO3 nanoparticles,” Solid State Commun.,
vol. 152, no. 6, pp. 525–529, 2012.
[166] A. Hussain et al., “The development of BiFeO3-based ceramics,” Chin.
Sci. Bull., vol. 59, no. 36, pp. 5161–5169, 2014.
[167] X. Chen et al., “Structure, ferroelectric and piezoelectric properties of
multiferroic Bi0. 875Sm0. 125FeO3 ceramics,” J. Alloys Compounds,
vol. 541, pp. 173–176, 2012.
[168] R. K. Mishra, D. K. Pradhan, R. N. P. Choudhary, and A. Baner-
jee, “Effect of yttrium on improvement of dielectric properties and
magnetic switching behavior in BiFeO3,” J. Phys., Condens. Matter,
vol. 20, no. 4, 2008, Art. no. 045218.
[169] X. Liang et al., “Soft magnetism, magnetostriction, and microwave
properties of Fe-Ga-C alloy films,” IEEE Magn. Lett., vol. 10, pp. 1–5,
Dec. 2018.
[170] X. C. Kan et al., “Magnetic/structural phase diagram and zero tem-
perature coefficient of resistivity in GaCFe3xCox (0x3.0),J.
Alloys Compounds, vol. 663, pp. 94–99, 2016.
YIFAN HE received the B.E. degree in electri-
cal engineering from Tianjin University, Tianjin,
China, in 2014, and the M.S. degree in electri-
cal and computer engineering from Northeastern
University, Boston, MA, USA, in 2016, where
he is currently working toward the Ph.D. degree
in electrical engineering. His research interests
include magnetic material and its application in
RF/microwave devices, energy harvesting devices.
BIN LUO received the B.E. degree in electri-
cal engineering from the Huazhong University of
Science and Technology, Wuhan, China, in 2020.
He is currently working toward the Ph.D. degree
in electrical engineering with Northeastern Uni-
versity, Boston, MA, USA. His research inter-
ests include gas discharge plasma and lightning
physics, electrical breakdown and lightning pro-
tection, computational chemical kinetics, fluid dy-
namics and electromagnetics, novel magnetic, fer-
roelectric, and multiferroic materials and their ap-
plications in RF/microwave electronics and spintronics, sensors and quantum
devices.
NIAN-XIANG SUN (Fellow, IEEE) received the
Ph.D. degree from Stanford University, Stanford,
CA, USA. He was a Scientist with IBM, Armonk,
NY, USA, and Hitachi Global Storage Technolo-
gies, San Jose, CA, USA. He is currently a Profes-
sor with the Electrical and Computer Engineering
Department and the Director of the W.M. Keck
Laboratory for Integrated Ferroics, Northeastern
University, Boston, MA, USA, and the Founder
and a Chief Technical Advisor of Winchester Tech-
nologies, LLC, Burlington, MA, USA. He has au-
thored or coauthored more than 280 publications. He holds more than 20
patents and patent applications. His research interests include novel magnetic,
ferroelectric, and multiferroic materials, devices, and subsystems. One of his
articles was selected as the Ten Most Outstanding Full Papers in the Past
Decade (2001–2010) in Advanced Functional Materials. Dr. Sun was the re-
cipient of the NSF CAREER Award, the ONR Young Investigator Award, and
the Søren Buus Outstanding Research Award. He is an editor of the Sensors
and IEEE TRANSACTIONS ON MAGNETICS and a Fellow of the Institute of
Physics and the Institution of Engineering and Technology. He has given more
than 180 plenary/keynote/invited presentations and seminars.
22 VOLUME 1, NO. 4, OCTOBER 2021
... On the other hand, non-topological ME materials are intensively investigated for the development of innovative applications in several areas, ranging from passive magnetic electronic to low-power spintronics [18]. In particular, there is an increasing interest in integrated magnetic devices, such as tunable inductors [19][20][21][22][23], actuators, or transformers [24][25][26]: these provide a new paradigm for circuit design of adaptive power converters (e.g. the transformer) or tunable multiband radiofrequency (RF) communications systems [2,21,27,28]. In this context, a natural question arises as to what extend TME materials are also suitable candidates to manufacture passive electronic components or magnetic devices endowed with certain desirable properties [14,[29][30][31][32]: a high inductance, a high operating frequency (e.g. 1 − 10 8 Hz) and a low power consumption. ...
... Notably, given a ME susceptibility χ ∼ 1000α 0 , the solenoid inductor composed of either the NI-TI or the TI-TI configuration manifests a self-inductance tunability of over 200% up to 100 GHz in the millimeter length scale (which corresponds to the yellow region in the middle and right upper panels of Fig. (5)). This is fairly comparable to previous proposals of ME voltage tunable inductors consisting of Metglas/PZT and nickel/cobalt ferrite composites, which have retrieved an inductance tunability around 50% [2,20,28] and 750% [4] up to 10 MHz, respectively (see also [22,23]). ...
... the amplitude of the axial magnetic field significantly grows by the azimuthal surface Hall current). This result supports the idea that ME materials in the near future could represent a promising platform for the implementation of integrable tunable inductors in the RF domain [21,28,40]. ...
Preprint
Full-text available
Despite the prospect of next-generation electronic technologies has spurred the investigation of the remarkable topological magnetoelectric response, it remains largely unexplored its potential in the application of basic electronic devices. In this paper, we undertake this task at the theoretical level by addressing the $\theta$-electrodynamics and examine electromagnetic properties (e.g. tunable inductance, operating frequency range, and power consumption) of three fundamental passive magnetic devices endowed with this effect: the primitive transformer, the bilayer solenoid inductor, and the solenoid actuator. We further exploit the methodology of magnetic circuits to obtain an extended Hopkinson's law that is valid for both topological and ordinary magnetoelectric responses (provided it is uniform in the bulk). Under low-power conditions, we find out that the functionally passive part of the topological-magnetoelectric transformer, solenoid inductor as well as solenoid actuator is indistinguishable from the conventional situation up to second-order in the magnetoelectric susceptibility; and argue that the main benefit of using topological insulators essentially relies on a lower power consumption. Our theoretical framework is also convenient to analyse magnetoelectric inductors endowed with a relatively large magnetoelectric susceptibility, they display a broad inductance tunability of over 250% up to 10 MHz. Conversely, our treatment predicts that the operating frequency range could be restricted below the ultra low frequency by a significantly strong magnetoelectric response (e.g. retrieved by certain multiferroic heterostructures).
... The use of electric fields for control of magnetism has been a long-term goal of magnetoelectronics 1 in its many manifestations, ranging from metal and semiconductor spintronics 2-4 to microwave electronics [5][6][7][8] to emerging applications in quantum information. 9 This interest arises from the potential for clear improvements in scaling, high-speed control, and multifunctional integration. ...
... (9) yields a magnetoelastic constant for V[TCNE]x of λs ∼ −1 ppm to λs ∼ −4 ppm for the devices measured here. This range shows excellent agreement with the DFT calculations of the magnetoelastic coefficientλ 100 = −2.52 ppm from Eq.(8). ...
Article
Full-text available
We demonstrate indirect electric-field control of ferromagnetic resonance (FMR) in devices that integrate the low-loss, molecule-based, room-temperature ferrimagnet vanadium tetracyanoethylene (V[TCNE]x∼2) mechanically coupled to PMN-PT piezoelectric transducers. Upon straining the V[TCNE]x films, the FMR frequency is tuned by more than 6 times the resonant linewidth with no change in Gilbert damping for samples with α = 6.5 × 10⁻⁵. We show this tuning effect is due to a strain-dependent magnetic anisotropy in the films and find the magnetoelastic coefficient |λs| ∼ (1–4.4) ppm, backed by theoretical predictions from density-functional theory calculations and magnetoelastic theory. Noting the rapidly expanding application space for strain-tuned FMR, we define a new metric for magnetostrictive materials, magnetostrictive agility, given by the ratio of the magnetoelastic coefficient to the FMR linewidth. This agility allows for a direct comparison between magnetostrictive materials in terms of their comparative efficacy for magnetoelectric applications requiring ultra-low loss magnetic resonance modulated by strain. With this metric, we show V[TCNE]x is competitive with other magnetostrictive materials, including YIG and Terfenol-D. This combination of ultra-narrow linewidth and magnetostriction, in a system that can be directly integrated into functional devices without requiring heterogeneous integration in a thin film geometry, promises unprecedented functionality for electric-field tuned microwave devices ranging from low-power, compact filters and circulators to emerging applications in quantum information science and technology.
... However, these approaches are slow, bulky and power-hungry and often introduce discrete tunable states, structural fragility. ME materials exhibit electric field and magnetic field control of high permeability and permittivity, ferromagnetic resonance (FMR), and non-uniform spin waves, which can be harnessed for tunable power, RF and microwave components, including inductors, filters, phase shifters and resonators 120 . The strain generated through piezoelectric or electrostrictive effects induces an effective magnetic field, altering magnetic properties and device performance 16,17 . ...
... The innovative design principles of ME antennas position them as promising solutions to the constraints faced by traditional antennas in VLF communication. Their ability to provide compact, lightweight and energy-efficient alternatives addresses the evolving demands of communication systems in challenging environments, making ME antennas a focal point of interest and research [3,16,17]. Due to the numerous advantages offered by ME antennas operating in VLF ranges, significant research efforts have recently been devoted to this area. In 2015, Yao et al. [18] proposed a multi-iron antenna based on bulk acoustic wave intermediation. ...
Article
Full-text available
VLF magneto-electric (ME) antennas have gained attention for their compact size and high radiation efficiency in lossy conductive environments. However, the need for a large DC magnetic field bias presents challenges for miniaturization, limiting portability. This study introduces a self-biased ME antenna with an asymmetric design using two magneto materials, inducing a magnetization grading effect that reduces the resonant frequency during bending. Operating principles are explored, and performance parameters, including the radiation mechanism, intensity and driving power, are experimentally assessed. Leveraging its excellent direct and converse magneto-electric effect, the antenna proves adept at serving as both a transmitter and a receiver. The results indicate that, at 2.09 mW and a frequency of 24.47 kHz, the antenna has the potential to achieve a 2.44 pT magnetic flux density at a 3 m distance. A custom modulation–demodulation circuit is employed, applying 2ASK and 2PSK to validate communication capability at baseband signals of 10 Hz and 100 Hz. This approach offers a practical strategy for the lightweight and compact design of VLF communication systems.
... Это, в свою очередь, дает возможность на их основе создавать датчики переменных и постоянных магнитных полей, автономные источники энергии, управляемые устройства электроники (индукторы и трансформаторы), антенны, новые типы магнитной памяти и др. [1][2][3]. МЭ эффекты в таких структурах возникают в результате комбинации магнитострикции ФМ слоя и пьезоэлектрического эффекта в ПЭ слое [4]. При помещении МЭ структуры во внешнее магнитное поле h ФМ слой деформируется вследствие магнитострикции. ...
Article
Full-text available
Objectives. The development of composite structures in which a strongly anisotropic magnetoelectric (ME) effect is observed is relevant for the creation of sensors that are sensitive to the direction of the magnetic field. Such an ME effect can arise due to the anisotropy of both the magnetic and the piezoelectric layers. In this work, a new anisotropic material named as a magnetostrictive fiber composite (MFC), comprising a set of nickel wires placed closely parallel to each other in one layer and immersed in a polymer matrix, is manufactured and studied. The study aimed to investigate the linear ME effect in a structure comprising of a new magnetic material, MFC, and lead zirconate titanate (PZT-19). Methods. The magnetostriction for the MFC structure was measured using the strain-gauge method; the ME effect was determined by low-frequency magnetic field modulation. Results . Structures with nickel wire diameters of 100, 150, and 200 μm were fabricated. The MFC magnetostriction field dependences were determined along with the frequency-, field-, and amplitude dependences of the ME voltage in the case of linear ME effect. Measurements were carried out at various values of the angle between the direction of the magnetic field and the wires. All samples demonstrated strong anisotropy with respect to the direction of the magnetic field. When the magnetic field orientation changes from parallel to perpendicular with respect to the nickel wire axes, the ME voltage decreases from its maximum value to zero. Conclusions. The largest ME coefficient 1.71 V/(Oe · cm) was obtained for a structure made of MFC with a wire diameter of 150 μm. With increasing wire diameter, the resonance frequency increases from 3.5 to 6.5 kHz. The magnetostriction of the MFC is comparable in magnitude to that of a nickel plate having the same thickness.
Article
Full-text available
Compact, conformal antennas with ground plane immunity and high gain are crucial to IoT, 5G, and biomedical applications. Conventional electrical antennas suffer large size, detuned impedance, and degraded gain and radiation efficiency on a ground plane as they operate on electromagnetic resonance and use electric dipole for radiation. Utilizing electromechanical resonance, magnetoelectric(ME) coupling, and magnetic dipole radiation in magnetostrictive/piezoelectric heterostructures, ME antennas exhibit ultra‐compact sizes comparable to acoustic wavelength and enhanced radiation performance on a ground plane. This study first utilizes parallel and series antenna array topology to achieve a profound gain and radiation efficiency enhancement without degrading impedance mismatch and quality factor of ME resonators. Notably, by increasing the array element number, 10 dBi gain enhancement is achieved in 3 × 3 thin‐film bulk acoustic resonator (FBAR) ME antenna array, reaching a peak antenna gain of −17.3 dBi. Unlike conventional antenna arrays, ME antenna arrays enhance radiation efficiency without affecting directivity, owing to ultra‐compact dimensions much less than one‐quarter of electromagnetic wavelength. Their ground plane immunity and 3 dB gain enhancement on ground planes with different shapes are also demonstrated. The demonstrated ME antenna arrays are outstanding platform‐independent ultra‐compact high‐gain conformal antenna candidates for wireless communication, wireless power transfer, and portable electronic and biomedical devices.
Article
Despite the prospect of next-generation electronic technologies spurring the investigation of the remarkable topological magnetoelectric response, its potential remains largely unexplored in the application of basic electronic devices. In this paper, we undertake this task at the theoretical level by addressing the θ electrodynamics and examine electromagnetic properties (e.g., tunable inductance, operating frequency range, and power consumption) of three fundamental passive magnetic devices endowed with this effect: the primitive transformer, the bilayer solenoid inductor, and the solenoid actuator. We further exploit the methodology of magnetic circuits to obtain an extended Hopkinson’s law that is valid for both topological and ordinary magnetoelectric responses (provided it is uniform in the bulk). Under low-power conditions, we find out that the functionally passive part of the topological-magnetoelectric transformer, the solenoid inductor as well as the solenoid actuator, is indistinguishable from the conventional situation up to second order in the magnetoelectric susceptibility; we argue that the main benefit of using topological insulators essentially relies on a lower power consumption. Our theoretical framework is also convenient to analyze magnetoelectric inductors endowed with a relatively large magnetoelectric susceptibility, they display a broad inductance tunability of over 200% up to 100 GHz in the millimeter length scale. Conversely, our treatment predicts that the operating frequency range could be restricted below the ultralow frequency by a significantly strong magnetoelectric response (e.g., retrieved by certain multiferroic heterostructures).
Article
Magnetoelectric (ME) antennas have offered significant advantages in the design of very-low-frequency (VLF) communication systems operating in challenging environments such as underground, underwater, and inside metallic enclosures. However, state of the art ME-based communication systems either depend on bulky commercial signal processing instruments or have limited reports on digital signal transmissions. In this work, we demonstrate a compact design of an ME VLF communication system based on self-designed modulator and demodulator circuits, working at the electromechanical resonance frequency of the ME antenna and having a remarkably reduced size for greater applicability in realistic industrial scenarios. The concept of sending Morse code through amplitude-shift-keying modulation was confirmed in the proposed ME VLF communication system. Our findings revealed that the maximum communication distance can be considered as 5.7 m when the bit-error-rate is limited to 10 <sup xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">-2</sup> . Furthermore, VLF wireless signal transfer through a harsh environment (metal box) was verified with our communication system. An effective communication distance of 0.85 m could be realized when placing the ME receiver in a 5 mm iron box, demonstrating the ability of the wireless signal to penetrate conductive media. The proposed fully packed ME VLF communication system provides advantageous alternative for underground, underwater, and inside-metal-containers communications.
Article
This issue of IEEE Microwave Magazine contains three focus articles covering nonreciprocal and tunable devices and circuits using novel magnetic materials, which provide essential functions for modern RF and microwave front ends. This multidisciplinary field involves materials, device physics, electromagnetics, circuit design, fabrication, and testing. Accordingly, microwave magnetics provides a rich landscape for scientific discovery and advanced device development with the potential of producing differentiating technologies that improve the performance of future RF and microwave systems.
Article
Full-text available
Parasitic magnetic noise arising from an electrical power transmission system is the most abundant form of waste energy in our daily life. In this work, a flexible and rollable magneto-mechano-electric nanogenerator (MMENG)-based wireless Internet of Things (IoT) sensor has been demonstrated to capture and utilize the environmental magnetic noise in the absence of a direct current magnetic field. Free-standing magnetoelectric composites are fabricated by combining magnetostrictive nickel ferrite (NiFe 2 O 4) nanoparticles (∼9 nm diameter) and piezoelectric polyvinylidene-co-trifluoroethylene polymer. The magneto-elctric 0−3-type nanocomposites possess maximum magnetoelectric voltage coefficient (α) of 11.43 mV/cm Oe. Even, without a magnetic bias field, 99% of the maximum α value is observed due to the self-bias effect. The magnetoelectric voltage generation capability under a low-frequency (50−60 Hz) alternating current magnetic field, validated by theoretical simulation, enables the nanocomposite to design efficient MMENG for harvesting a low-frequency stray magnetic field from the power cable of home appliances, such as electric kettle and microwave oven. As a result, the MMENG generates a peak-to-peak open circuit voltage of 1.4 V and output power density of 0.05 μW/cm 3 and successfully operates a commercial capacitor under the weak (∼1.7 × 10 −3 T) and low-frequency (∼50 Hz) stray magnetic field arising from the power cable of electric kettle. Additionally, under the rolled condition around the power cable, the MMENG generates a slightly improved peak-to-peak open circuit voltage of 1.5 V. Finally, the harvested electrical signal has been wirelessly transmitted to a smart phone to demonstrate the possibility of position monitoring system construction. This cost-effective and easy to integrate approach with tailored size and shape of device configuration is expected to be explored in next-generation self-powered IoT sensors including implantable biomedical devices and human health monitoring sensory systems.
Article
Full-text available
Many applications require galvanic isolation between the circuit where the current is flowing and the measurement device. While for AC, the current transformer is the method of choice, in DC and, especially for low currents, other sensing methods must be used. This paper aims to provide a practical method of improving the sensitivity and linearity of a giant magnetoresistance (GMR)-based current sensor by adapting a set of design rules and methods easy to be implemented. Our approach utilizes a multi-trace current trace and a double differential GMR based detection system. This essentially constitutes a planar coil which would effectively increase the usable magnetic field detected by the GMR sensor. An analytical model is developed for calculating the magnetic field generated by the current in the GMR sensing area which showed a significant increase in sensitivity up to 13 times compared with a single biased sensor. The experimental setup can measure both DC and AC currents between 2–300 mA, with a sensitivity between 15.62 to 23.19 mV/mA, for biasing fields between 4 to 8 Oe with a detection limit of 100 μA in DC and 100 to 300 μA in AC from 10 Hz to 50 kHz. Because of the double differential setup, the detection system has a high immunity to external magnetic fields and a temperature drift of the offset of about −2.59 × 10−4 A/°C. Finally, this setup was adapted for detection of magnetic nanoparticles (MNPs) which can be used to label biomolecules in lab-on-a-chip applications and preliminary results are reported.
Article
Full-text available
Thanks to high sensitivity, excellent scalability, and low power consumption, magnetic tunnel junction (MTJ)-based tunnel magnetoresistance (TMR) sensors have been widely implemented in various industrial fields. In nondestructive magnetic flux leakage testing, the magnetic sensor plays a significant role in the detection results. As highly sensitive sensors, integrated MTJs can suppress frequency-dependent noise and thereby decrease detectivity; therefore, serial MTJ-based sensors allow for the design of high-performance sensors to measure variations in magnetic fields. In the present work, we fabricated serial MTJ-based TMR sensors and connected them to a full Wheatstone bridge circuit. Because noise power can be suppressed by using bridge configuration, the TMR sensor with Wheatstone bridge configuration showed low noise spectral density (0.19 μV/Hz0.5) and excellent detectivity (5.29 × 10−8 Oe/Hz0.5) at a frequency of 1 Hz. Furthermore, in magnetic flux leakage testing, compared with one TMR sensor, the Wheatstone bridge TMR sensors provided a higher signal-to-noise ratio for inspection of a steel bar. The one TMR sensor system could provide a high defect signal due to its high sensitivity at low lift-off (4 cm). However, as a result of its excellent detectivity, the full Wheatstone bridge-based TMR sensor detected the defect even at high lift-off (20 cm). This suggests that the developed TMR sensor provides excellent detectivity, detecting weak field changes in magnetic flux leakage testing.
Article
Full-text available
Nanomagnets form the building blocks for a gamut of miniaturized energy-efficient devices including data storage, memory, wave-based computing, sensors, and biomedical devices. They also offer a span of exotic phenomena and stern challenges. The rapid advancements of nanofabrication, characterization, and numerical simulations during the last two decades have made it possible to explore a plethora of science and technology applications related to nanomagnet dynamics. The progress in the magnetization dynamics of single nanomagnets and one- and two-dimensional arrays of nanostructures in the form of nanowires, nanodots, antidots, nanoparticles, binary and bi-component structures, and patterned multilayers have been presented in detail. Progress in unconventional and new structures like artificial spin ice and three-dimensional nanomagnets and spin textures like domain walls, vortex, and skyrmions has been presented. Furthermore, a huge variety of new topics in the magnetization dynamics of magnetic nanostructures are rapidly emerging. A future perspective on the steadily evolving topics like spatiotemporal imaging of fast dynamics of nanostructures, dynamics of spin textures, and artificial spin ice have been discussed. In addition, dynamics of contemporary and newly transpired magnetic architectures such as nanomagnet arrays with complex basis and symmetry, magnonic quasicrystals, fractals, defect structures, and novel three-dimensional structures have been introduced. Effects of various spin–orbit coupling and ensuing spin textures as well as quantum hybrid systems comprising of magnon–photon, magnon–phonon, and magnon–magnon coupling and antiferromagnetic nanostructures have been included. Finally, associated topics like nutation dynamics and nanomagnet antenna are briefly discussed. Despite showing great progress, only a small fraction of nanomagnetism and its ancillary topics have been explored so far and huge efforts are envisaged in this evergrowing research area in the generations to come.
Article
Full-text available
Since the revival of multiferroic laminates with giant magnetoelectric (ME) coefficients, a variety of multifunctional ME devices, such as sensor, inductor, filter, antenna etc. have been developed. Magnetoelastic materials, which couple the magnetization and strain together, have recently attracted ever-increasing attention due to their key roles in ME applications. This review starts with a brief introduction to the early research efforts in the field of multiferroic materials and moves to the recent work on magnetoelectric coupling and their applications based on both bulk and thin-film materials. This is followed by sections summarizing historical works and solving the challenges specific to the fabrication and characterization of magnetoelastic materials with large magnetostriction constants. After presenting the magnetostrictive thin films and their static and dynamic properties, we review micro-electromechanical systems (MEMS) and bulk devices utilizing ME effect. Finally, some open questions and future application directions where the community could head for magnetoelastic materials will be discussed.
Article
In order to advance the potential of thick on-chip hard magnets for the micro-electro-mechanical system (MEMS), we investigate a new silicon molding technique to fabricate dry-packed NdFeB magnets, including a silicon compression tool, which enables the pressing step during silicon-compatible processing. This process delivers samples with a remanence of 0.42 T and an energy product of 38 kJ/m <sup xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">3</sup> . Further studies of metal molding show that, for wax-bonding powder-based NdFeB magnets, the optimum fabrication condition is 300 °C and 425 MPa, giving a remanence of 0.54 T and an energy product of 61.7 kJ/m <sup xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">3</sup> .
Article
In this paper, a highly sensitive and specific technique based on the principle of giant magnetoresistance (GMR) has been proposed for the early stage Tuberculosis (TB) diagnostics. This GMR biosensing assay employs monoclonal antibodies against M. tuberculosis specific ESAT-6 antigen with the use of magnetic nanoparticles (MNPs) as labels. MNPs bind to the GMR sensor in presence of ESAT-6 and the binding is proportional to the ESAT-6 protein concentration leading to the change in overall resistance of GMR sensor. GMR biosensor simulation showed that ESAT-6 concentration can be detected in the range of pg/mL in comparison to the other transduction techniques available for ESAT-6 detection and further, the signal strength increased with the increase in the concentration. This work has shown that the GMR biosensing strategy is pertinent for the TB detection at the primitive phases when compared with other magnetic techniques used for TB diagnostics.
Article
Ferroelectric switching is unambiguously demonstrated for the first time in a III-V semiconductor based material: Al1-xScxN—A discovery which could help to satisfy the urgent demand for thin film ferroelectrics with high performance and good technological compatibility with generic semiconductor technology which arises from a multitude of memory, micro/nano-actuator, and emerging applications based on controlling electrical polarization. The appearance of ferroelectricity in Al1-xScxN can be related to the continuous distortion of the original wurtzite-type crystal structure towards a layered-hexagonal structure with increasing Sc content and tensile strain, which is expected to be extendable to other III-nitride based solid solutions. Coercive fields which are systematically adjustable by more than 3 MV/cm, high remnant polarizations in excess of 100 μC/cm2—which constitute the first experimental estimate of the previously inaccessible spontaneous polarization in a III-nitride based material, an almost ideally square-like hysteresis resulting in excellent piezoelectric linearity over a wide strain interval from −0.3% to + 0.4% and a paraelectric transition temperature in excess of 600 °C are confirmed. This intriguing combination of properties is to our knowledge as of now unprecedented in the field of polycrystalline ferroelectric thin films and promises to significantly advance the commencing integration of ferroelectric functionality to micro- and nanotechnology, while at the same time providing substantial insight to one of the central open questions of the III-nitride semiconductors—that of their spontaneous polarization.
Article
MEMS sensors based on magnetoelectric composites have attracted great interest due to their capability to detect weak magnetic fields, showing a high potential in applications like biomagnetic field detection and magnetic particle imaging. This paper reports on a scandium aluminum nitride thin film based MEMS magnetoelectric sensor. The sensor consists of a polycrystalline silicon cantilever with a size of 1000 µm x 200 µm covered by a piezoelectric Al0.73Sc0.27N and a magnetostrictive (Fe90Co10)78Si12B10 thin film. The performance of the presented sensor is investigated based on ME voltage coefficient, voltage noise density and limit of detection, and compared to the characteristics of aluminum nitride thin film based ME sensor with the same layout and fabrication technology. By using Al0.73Sc0.27N thin film with a higher piezoelectric activity instead of AlN in MEMS ME sensors, ME voltage coefficient of (1334 ± 84) V/cmOe in resonance is almost double, thereby lowering the requirements for the electronic system. The limit of detection of (60 ± 2) pT/Hz0.5 remains unchanged due to the dominant thermomechanical noise in resonance.