77
77
sciences
Article
Analytical Electromechanical Modeling of Nanoscale
Flexoelectric Energy Harvesting
Yaxuan Su 1 , Xiaohui Lin 2 , Rui Huang 3 and Zhidong Zhou 2,4, *
1 Chengyi University College, Jimei University, Xiamen 361021, China; [email protected]
2 Department of Materials Science and Engineering, College of Materials, Xiamen University,
Xiamen 361005, China; [email protected]
3 Department of Aerospace Engineering and Engineering Mechanics, University of Texas,
Austin, TX 78712, USA; [email protected]
4 Fujian Provincial Key Laboratory of Advanced Materials, Xiamen University, Xiamen 361005, China
* Correspondence: [email protected]
Received: 30 April 2019; Accepted: 31 May 2019; Published: 1 June 2019
Abstract: With the attention focused on harvesting energy from the ambient environment for
nanoscale electronic devices, electromechanical coupling effects in materials have been studied for
many potential applications. Flexoelectricity can be observed in all dielectric materials, coupling the
strain gradients and polarization, and may lead to strong size-dependent effects at the nanoscale.
This paper investigates the flexoelectric energy harvesting under the harmonic mechanical excitation,
based on a model similar to the classical Euler–Bernoulli beam theory. The electric Gibbs free energy
and the generalized Hamilton’s variational principle for a flexoelectric body are used to derive the
coupled governing equations for flexoelectric beams. The closed-form electromechanical expressions
are obtained for the steady-state response to the harmonic mechanical excitation in the flexoelectric
cantilever beams. The results show that the voltage output, power density, and mechanical vibration
response exhibit significant scale effects at the nanoscale. Especially, the output power density for
energy harvesting has an optimal value at an intrinsic length scale. This intrinsic length is proportional
to the material flexoelectric coefficient. Moreover, it is found that the optimal load resistance for peak
power density depends on the beam thickness at the small scale with a critical thickness. Our research
indicates that flexoelectric energy harvesting could be a valid alternative to piezoelectric energy
harvesting at micro- or nanoscales.
1. Introduction
With the development of nanotechnology, harvesting ambient waste energy into usable energy
has drawn growing attention in the last decades. One of the aims in this field is to provide power
for small electronic devices by harvesting ambient energy [1]. As stated by Williams and Yates [2] in
their early work on harvesting ambient waste vibrational energy for microsystems, there are three
basic vibration-to-electric energy conversion mechanisms: electromagnetic [2,3], electrostatic [4,5],
and piezoelectric transduction [6,7]. Using piezoelectric nanomaterials as ambient energy harvesting is
considered as a promising way to supply normal microelectronic devices, such as environmental or
biomedical devices, portable multimedia, distributed sensor networks, or mobile communication [8].
In addition, some important system-level and circuit-level works about micro-scale energy harvesting
systems have been reported [9–11].
Piezoelectricity, which generally assumes a linear relationship between electric field and strain,
exists only in non-centrosymmetric dielectric materials. Alternatively, flexoelectricity, which is the
coupling of electrical polarization and the strain gradient, exists in a wide variety of dielectric materials
and may lead to strong size-dependent properties at the nanoscale. A series of experimental [12–14]
and theoretical works [15–17] for the flexoelectric effect have been reported. Flexoelectric effects
may be exploited to enhance piezoelectric properties of materials [18,19], or to enable new classes
of electromechanically-coupled materials, such as lead-free “pseudo-piezoelectric” [16]. In addition,
one can exploit flexoelectricity to construct materials of non-uniform shapes, which exhibit large
strain gradients and can generate electricity despite being non-piezoelectric [20]. Recently, several
comprehensive reviews on flexoelectricity of solid crystals, thin films, polymers, and living membranes
have been published [21–23].
In the nano-electromechanical systems (NEMSs), such as resonators, sensors, actuators, and energy
harvesters, the electromechanical coupling of the nanobeams and nanoplates with the flexoelectric
effect has drawn a surge of interest due to the relatively large flexoelectric effect at the nanoscale.
The electromechanical coupling responses of the nanobeams and nanoplates have been analyzed by
numerical and analytical methods for static and dynamic problems [17,24–29]. Based on the linear
piezoelectricity theory developed by Toupin [30], Shen and Hu [31] have established a theoretical
framework by a variational principle for dielectrics including the electrostatic force, flexoelectricity,
and surface effects. Based on the electric Gibbs energy, Liang et al. [32] proposed the Euler–Bernoulli
beam model to investigate the effect of surface and flexoelectricity on the coupling response of
piezoelectric nanostructures. They found that the effective bending rigidity of the nanobeam enhances
dramatically in nanoscale. Yan and Jiang [33] discussed the flexoelectric effect on the mechanical
and electrical properties of piezoelectric nanobeams under static bending and different mechanical
boundary conditions based on the internal energy. Simulation results show that the flexoelectric
effect is sensitive to the mechanical boundary conditions and the direction of applied electric field.
Liang et al. [34] investigated the effect of flexoelectricity and surface on buckling and vibration behavior
of piezoelectric beams by use of the Euler–Bernoulli beam model. They found that the effects of surface
and flexoelectricity influenced the resonance frequency of piezoelectric nanowires. Based on the
Kirchhoff plate model and the extended linear piezoelectric theory, Wang et al. [35] presented a finite
difference method to solve the non-conventional governing equations of the cantilevered piezoelectric
nanoplates. Recently, Zhou et al. [36] have investigated the flexoelectric effect in piezoelectric
nanobeams with three different electrical boundary conditions. The induced electric potential due to
the flexoelectric effect has been obtained under the open circuit conditions, which may be important
for sensing or energy harvesting applications. For the flexoelectric energy harvesting applications,
Deng et al. [37] discussed the flexoelectric energy harvester in the nanoscale. Based on the internal
energy density, they obtained governing equations and solved the frequency response functions using
the assumed-modes method. Moura and Erturk [38] applied the distributed-parameter method to
discuss the flexoelectric energy harvesters in elastic dielectrics. Considering the surface effects, Yan [39]
analyzed the flexoelectric energy harvest by use of the assumed-modes method. Liang et al. [40]
developed flexoelectric-piezoelectric energy harvesters based on the Timoshenko laminated beams
model. The three-layered energy harvesters in parallel and series configurations have been discussed in
detail. However, the closed-form analytical solution for the flexoelectric-piezoelectric energy harvesting
has not been reported.
In this paper, we focus on the flexoelectric energy harvesting system, which is a traditional
cantilever piezoelectric beam model with a tip mass, by the coupled distributed-parameter model.
Based on the electric Gibbs free energy [36] and generalized Hamilton’s variational principle,
the electromechanically-coupled dynamic and electrical circuit equations with the flexoelectric effect are
derived. The closed-form analytical solution is presented for the flexoelectric energy harvesting under
base excitations. The electromechanical frequency response functions (FRFs) that relate the voltage
output, power density, and mechanical vibration response are derived for harmonic excitations in
Appl. Sci. 2019, 9, 2273 3 of 18
closed-form.
Appl. Sci. 2019, 9,Based
x FOR onPEERtheREVIEW
multi-mode solutions, the performance of the flexoelectric energy harvesting 3 of 18
is analyzed in detail.
2. Electromechanical System and Mathematical Formation
2. Electromechanical System and Mathematical Formation
The present study focuses on the vibration responses of a flexoelectric cantilever beam with
lengthThe 𝐿, present
width 𝐵, study focuses
thickness ℎ, and
on thea tip mass 𝑀 responses
vibration as shown of a flexoelectric
in Figure cantilever
1, in which the beam beam with
is coated
length L, width B, thickness h, and
by conductive electrodes on both upper and lower a tip mass M t as shown in Figure 1, in which the beam
surfaces. We assumed the electrode layers wereis coated by
conductive
thin so thatelectrodes
the effecton ofboth
theirupper andon
stiffness lower surfaces. responses
mechanical We assumed of the electrode
system was layers were thinThe
negligible. so
that the effect
coordinate of their
system , 𝑥 , 𝑥 ) ison
(𝑥 stiffness mechanical
shown responses
in Figure of the system
1. The motion of the was
basenegligible. The coordinate
for the flexoelectric beam
system
was in the (x1 , 𝑥x2 , direction.
x3 ) is shown With in aFigure 1. The
transverse motion
base of the 𝑤
excitation base(𝑡),
forthe
thecantilever
flexoelectric
beambeam was in the
vibrates
xbending
3 direction.mode. WithAsa atransverse
result of basethe dynamic wb (t)gradient
excitationstrain , the cantilever
causedbeam vibrates in
by vibration the bending
responses, the
mode. As a result of the dynamic strain gradient caused by vibration
flexoelectric beam produces an alternating potential difference across the surface electrodes. These responses, the flexoelectric
beam produces
electrodes werean alternating
connected to potential
an external difference
resistanceacross
(R) the surface electrodes.
to quantify the electricThese electrodes
potential were
and power
connected
output. Here, to anthe external
internalresistance
electrical(R) resistance
to quantifyofthe theelectric potentialbeam
flexoelectric and power
was not output.
takenHere,into
the internal electrical resistance of the flexoelectric beam was not taken into
consideration since it can be regarded as a resistance connected in parallel to the external electrical consideration since it can
be regarded
load [37,41]. as a resistance connected in parallel to the external electrical load [37,41].
Figure 1. AAcantilever
Figure cantileverbeam
beammodel
modelof of
thethe flexoelectric
flexoelectric energy
energy harvesting
harvesting withwith
a tipamass
tip mass under
under base
base excitations.
excitations.
The
The generalized
generalized Hamilton’s
Hamilton’s variational
variational principle
principle for
for flexoelectric
flexoelectric materials
materials can
can be
be written
written as
as [32]
[32]
δ [𝐾 − 𝐺 + 𝑊]d𝑡 = 0,
Z T
(1)
δ [K − G + W ]dt = 0, (1)
where 𝐾, 𝐺, and 𝑊 are the total kinetic 0energy, total electric Gibbs free energy, and external work,
respectively.
where In this
K, G, and study,
W are the the
totaltip mass energy,
kinetic was regarded as a particle
total electric Gibbs and
free the rotary
energy, andinertia of the
external tip
work,
mass was notInconsidered,
respectively. which
this study, the has less
tip mass was influence.
regarded asAssuming
a particlenoandexternal body
the rotary forces
inertia andtip
of the electric
mass
field,not
was forconsidered,
the flexoelectric cantilever
which has less beam showing
influence. in Figure
Assuming no1,external
Equation (1) could
body forcesbe
andrewritten
electric as
field,
for the flexoelectric cantilever beam showing in Figure 1, Equation (1) could be rewritten as
δ d𝑡 [ 𝜌|𝑤 | − 𝑔 ] d𝑉 + δ 𝑀 |𝑤 | d𝑡 | − d𝑡 ∮ 𝜛δ𝜓d𝐴 = 0, (2)
Z T Z Z T Z T I
of1 ρthis 1 . m 2 (𝑥 , 𝑡) = 𝑤 (𝑡) + 𝑤(𝑥 , 𝑡) is the absolute
where 𝜌 is δ density
dt
.m2
w flexoelectric
− g1 dV + δ material, Mt w𝑤 dt x1 =L − dt $δψdA = 0, (2)
displacement in 0 the V 𝑥 2direction, 𝑤 = 𝑤(𝑥 , 𝑡) 0 is2 the transverse displacement
0 s (relative to the base)
of the neutral surface at point 𝑥 , and 𝑔 is the general electric Gibbs free energy density. 𝜛(𝑡) and
where ρ is density of this flexoelectric material, wm (x1 , t) = wb (t) + w(x1 , t) is the absolute displacement
𝜓(𝑡) are the electric charge density and electric potential on the surface electrodes, which form the
in the x3 direction, w = w(x1 , t) is the transverse displacement (relative to the base) of the neutral
virtual work due to the moving charges on or out of the electrodes. It should be noted that the external
surface at point x1 , and g1 is the general electric Gibbs free energy density. $(t) and ψ(t) are the electric
work done by the surface hyper-stress tractions was ignored in this case.
charge density and electric potential on the surface electrodes, which form the virtual work due to
For the flexoelectric material, the general electric Gibbs free energy density function 𝑔 can be
the moving charges on or out of the electrodes. It should be noted that the external work done by the
expressed as [25,36]
surface hyper-stress tractions was ignored in this case.
For the flexoelectric material, 𝑔the=general
𝜎 𝜀 + 𝜎 𝜀Gibbs
electric , − free
𝐷 𝐸 ,energy density function g1 can(3)
be
expressed as [25,36]
where 𝜎 is the classical Cauchy stress 1tensor, 𝜀 1 is the strain 1 tensor, 𝜎 is the higher-order stress
g = σ
tensor, 𝐷 is the electric displacement 2vector, andε + σ ε − Di Ei , (3)
2 𝐸 is the2 electric field vector. In this paper, the
1 ij kl ijk ij,k
kinematics of the classical Euler–Bernoulli beam model was adopted to analyze the bending vibration
of the flexoelectric beam for energy harvesting. The relative displacement field 𝒖 in the Euler–
Bernoulli model is
Appl. Sci. 2019, 9, 2273 4 of 18
where σij is the classical Cauchy stress tensor, εij is the strain tensor, σijk is the higher-order stress tensor,
Di is the electric displacement vector, and Ei is the electric field vector. In this paper, the kinematics
of the classical Euler–Bernoulli beam model was adopted to analyze the bending vibration of the
flexoelectric beam for energy harvesting. The relative displacement field u in the Euler–Bernoulli
model is
∂w
( )
u = −x3 , 0, w , (4)
∂x1
where the displacement in the x2 direction is set to be zero as in the plane strain elasticity. With Equation (4)
for infinitesimal deformation, the only non-zero strain component and its gradients are
∂2 w ∂2 w ∂3 w
ε11 = −x3 , ε 11,3 = − , ε11,1 = −x3 , (5)
∂x21 ∂x21 ∂x31
where the stain gradient ε11,3 is essentially the bending curvature of the beam and ε11,1 is proportional
to the gradient of curvature. The latter is typically small compared to the former in the Euler–Bernoulli
beam model and may be neglected for a slender beam. For the same reason, only one component of
the electric field, E3 , was considered due to the relatively small electric field E1 in the length direction.
According to the constitutive equations of the flexoelectric beams, the non-zero stress, higher-order
stress, and electric displacement are [25,32,36]
where σ113 is defined as the higher order stress or the moment stress [25,36], which is induced by the
electric field due to the flexoelectric effect, c1111 is an elastic modulus, e311 is a piezoelectric coefficient,
µ3113 is a flexoelectric coefficient, and κ33 is a dielectric coefficient.
In the absence of free body charges, Gauss’s law of electrostatics leads to
∂2 Φ e
= 311 ε11,3 , (7)
∂x23 κ33
where Φ = Φ (x1 , x3 , t) is the electric potential of the beam and is related to the electric field by
∂Φ
E3 = − ∂x . Considering the surface electrical boundary conditions Φ x1 , x3 = 2h , t = ψ1 (x1 , t) and
3
Φ x1 , x3 = − h2 , t = ψ2 (x1 , t), the dynamic electric potential Φ and dynamic electric field E3 can be
obtained from Equation (7) as
e311 ∂2 w h2
!
1 x3
2
Φ(x1 , x3 , t) = − x3 + + ψ(x1 , t) + ψ2 (x1 , t), (8)
2κ33 ∂x2 4 2 h
1
e311 ψ ( x1 , t )
E3 = − ε11 − , (9)
κ33 h
where ψ(x1 , t) = ψ1 (x1 , t) − ψ2 (x1 , t) is the potential difference or voltage between the both surface
electrodes. Substituting Equation (9) into Equation (6), the electric displacement, stress, and higher
order stress could be expressed as
ψ
D3 = µ3113 ε11,3 − κ33 h
2
ψ
e
σ11 = c1111 + κ311 ε11 + e311 h .
(10)
33
σ = µ3113 e311 ε + µ ψ
113 κ33 11 3113 h
Appl. Sci. 2019, 9, 2273 5 of 18
Then, by Equation (3), the electric Gibbs free energy density is obtained as follows [36]:
e2311 2 e µ ψ ψ2
1 ε + 311 3113 ε11 ε11,3 + µ3113 ε11,3 − 1 κ33 .
g1 = c1111 + (11)
2 κ33 11 κ33 h 2 h2
Using Equations (5) and (6), the variational expression of general electric Gibbs free energy can be
written as [36]
R T R L ∂2 ψ ψB
RT R
δ 0 dt V g1 dV = 0 0 Gp ∂ w4 − µ3113 B 2 δw − µ3113 B ∂ w2 + κ33 h δψ dx1 dt+
4 2
∂x ∂x ∂x1
R T 1 1 (12)
∂2 w ∂w
R T ∂3 w
G p 2 − µ3113 Bψ δ ∂x x1 =L
dt − G p 3 δw x1 = L dt,
0 ∂x1 0 1 ∂x1
e2311
Bh3
where Gp = 12 c1111 + κ33 is the effective bending rigidity of the piezoelectric beam. It should be
noted that the present effective bending rigidity is quite different from the previous result in [37], where
the effective bending rigidity was derived from an internal energy density function. The effective
bending rigidity in [37] depends on the flexoelectricity of the material and becomes negative when the
beam thickness is small (~several nanometers). In contrast, the effective bending rigidity Gp in the
present model remains positive, independent of the flexoelectric coefficients.
Substituting Equation (12) into Equation (2), the generalized Hamilton’s variational equation of
the flexoelectric beam can be rewritten as
R T R L .. .. ..
R
T R L
Gp ∂ w4 −
4
dt Bh ρ w + w b δwdx 1 + M w
t b δ ( x1 − L ) δw + dt
0 0 0 0 ∂x1
∂2 ψ ψ
R T
µ3113 B 2 δw − µ3113 B ∂ w2 + κ33 B h δψ dx1 dt+ 0 Gp ∂ w2 −
2 2
∂x1 ∂x1 ∂x1 (13)
R T RT RL
∂w ∂ w
3 ∂ w
2
µ3113 Bψ δ ∂x x1 =L dt − 0 Gp 3 − Mt ∂t2 δw x1 =L dt − 0 dt 0 B$δψdx1 dt =
1 ∂x1
0,
where δ(x1 ) is the Dirac delta function. In deducing Equation (13), the following result has been used
Z T Z T Z T
1 .m2 .. ..
δ Mt w dt x1 =L = − Mt wb δ(x1 − L)δwdt − Mt w x1 =L δwdt. (14)
0 2 0 0
∂4 w ∂2 ψ ∂ 2 wb ∂2 w
Gp − µ3113 B + [m + Mt δ(x1 − L)] + m = 0, (15)
∂x41 ∂x21 ∂t2 ∂t2
L
d2 w ψ
Z
(µ3113 2
+ κ33 + $)dx1 = 0. (16)
0 dx1 h
When the external loading resistor R is connected into the surface electrodes, the electric current
ψ R .
R must be equal to the time rate of change of the average output positive charges, i.e., − 1h V $dV.
Using Equation (16), from Gauss’s law [37,41,42], the electromechanically-coupled electrical circuit
equation with flexoelectric effect can be written as
Appl. Sci. 2019, 9, 2273 6 of 18
.
ψ ψ ∂2 w
κ33 BL + = −µ3113 B , (17)
h R ∂x1 ∂t x1 = L
where ψ is assumed to be independent of x1 . In addition, the boundary conditions are obtained from
Equation (13) as
G ∂2 w
= µ3113 Bψ
p ∂x2
1 x1 = L
. (18)
Gp ∂∂xw3 = Mt ∂∂tw2
3 2
x1 = L
1 x1 =L
Apparently, the flexoelectric effect induces an effective bending moment at the end of the beam,
whereas the inertia of the tip mass induces a shear force at the end.
∂4 w ∂2 ψ ∂5 w ∂w ∂2 w ∂2 w
Gp − µ3113 B + cs + ca + [m + Mt δ(x1 − L)] 2b + m 2 = 0, (19)
∂x41 ∂x21 ∂x41 ∂t ∂t ∂t ∂t
where ca is the viscous air damping coefficient and cs is the strain-rate damping coefficient [42].
Viscous air damping is a simple way of modeling the force acting on the beam due to the air particles
displaced during vibration, while strain-rate damping accounts for the structural damping due to
energy dissipation of the beam material. Both of them satisfy the proportional damping criterion and
they are mathematically convenient for the modal analysis [42].
Based on the proportional damping assumption, the vibration response of the flexoelectric
cantilever beam can be represented as an absolutely and uniformly convergent series of eigenfunctions as
∞
X
w ( x1 , t ) = ∅r (x1 )ηr (t), (20)
r=1
where ∅r (x) is the mass normalized eigenfunction of the rth vibrational mode, which satisfies the
mechanical and the short circuit electrical conditions, and ηr (t) is the modal mechanical response
coefficient in the modal coordinates. For the piezoelectric or flexoelectric Euler–Bernoulli cantilever
beams, the eigenfunctions ∅r (x) can be obtained from the corresponding undamped free vibration
problem [41]
βr βr βr βr
" !#
∅r (x1 ) = Cr cos x1 − cos h x1 + ζr sin x1 − sin h x1 , (21)
L L L L
where ζr is
Mt
sin βr − sin hβr + βr mL (cos βr − sin hβr )
ζr = Mt
, (22)
cos βr + cos hβr − βr mL (sin βr − sin hβr )
and βr is the rth root of the transcendental characteristic equation
Mt
1 + cos βr cos hβr + βr (cos βr sin hβr − sin βr cos hβr ) = 0. (23)
mL
Appl. Sci. 2019, 9, 2273 7 of 18
In Equation (21), the constant Cr is the modal amplitude constant, which could be solved by
normalizing the eigenfunctions using the following orthogonality condition [33]
Z L
∅r (x1 )m∅s (x1 )dx1 + ∅r (L)Mt ∅s (L) = δrs , (24)
0
where δrs is Kronecker delta, defined as being equal to unity for s = r and equal to zero for s , r.
It should be noted that the undamped natural frequency of the rth vibration mode in the short circuit
condition is r
2
Gp
ωr = βr . (25)
mL4
It is important to mention that the electric potential difference between the two surface electrodes
due to the flexoelectric effect is a function of time t and as a result, the spatial derivative of ψ would
vanish in Equation (19). To include the flexoelectric effect in the governing Equation (19), Erturk and
Inman [41] assumed that ψ = ψ(x1 , t) = V (t)[H (x1 ) − H (x1 − L)], where V (t) is the output voltage and
H (x) is the Heaviside function.
To investigate the dynamic response of the system, the mode-superposition method was adopted
to solve the damped flexoelectric cantilever beam with the tip mass. Substituting the modified electric
potential expression and Equation (20) into Equation (19) and using the orthogonality condition (24),
the electromechanical coupling dynamic governing equation of the damped flexoelectric beams in
modal coordinates can be written as
d2 ηr (t)
2 η (t) + dηr (t) 2ξ ω = L µ dδ(x1 ) dδ(x1 −L)
R
dt2
+ ωr r r r 3113 BV ( t )∅ (
r 1x ) − dx1 −
dt
2
0 2
dx1 dx1
(26)
d w (t) L d wb ( t )
R
m dtb2 0
∅ r ( x 1 ) dx 1 − M t ∅ r ( L ) dt2 ,
where ξr is defined as the modal damping ratio of rth which include the viscous air damping part
ca and the strain-rate damping part cs [37,43]. Commonly, in experimental modal analysis practice,
one can identify the modal damping ratio ξr directly from the frequency response or time-domain
measurements, which avoids the requirement of defining and obtaining the physical damping terms cs
and ca [41]. Here, the nth derivative of the Dirac delta function satisfies the condition as follows
n
dn δ(x − x0 ) n d γ(x)
Z +∞
γ ( x ) dx = ( −1 ) x = x0 . (27)
−∞ dxn dxn
where fr (t) is the modal mechanical force function corresponding to the base excitation
d2 wb (t) L
d2 wb (t)
" Z #
fr (t) = −m ∅r (x1 )dx1 − Mt ∅r (L) . (29)
dt2 0 dt2
Similarly, substituting Equation (20) into Equation (17), the modal circuit equation can be
obtained as
∞
κ33 BL dV (t) V (t) X dηr (t)
+ =− Kr , (30)
h dt R dt
r=1
d∅r (x)
where Kr = µ3113 B dx .
x1 =L
Appl. Sci. 2019, 9, 2273 8 of 18
With the small deformation and linear system assumption, the modal mechanical and electrical
responses were in steady state and also harmonic with the same frequency as the base excitation. Thus,
the modal mechanical response and voltage response can be written as ηr (t) = Hr ejωt and V (t) = vejωt ,
respectively, in which both Hr and v are generally complex valued. Then, Equations (28) and (30)
lead to h 2 i
ωr − ω2 + 2ξr ωr jω Hr − vKr = Fr ,
∞ (32)
κ33 BL
1 P
h jω + R v + jω Kr Hr = 0.
r=1
Solving Equation (32), Hr and v are obtained explicitly. The analytical modal expressions of
voltage V and mechanical response ηr (t) can be obtained as follows
P∞ Kr Fr
jω r=1
(ω2r −ω2 )+2jωξr ωr
V (t) = − ejωt , (33)
κ33 BL jωKr2
P∞
1
R + h jω + r=1 (ω2 −ω2 )+2jωξr ωr
r
Fr Kr
jωt
ηr (t) = + V (t) e . (34)
ω2 − ω2 + 2jωξ ω ω − ω + 2jωξ ω
2 2
r r r r r r
We can then substitute ηr (t) into Equation (20) to obtain the deflection response w(x1 , t).
-2
10
(a) R1=1MΩ
R2=5MΩ
R3=10MΩ
R4=50MΩ
-4 R5=100MΩ
10 R6=500MΩ
|Voltage FRFs| [V/g]
Open-circuit
-6
10
-8
10
-10
10
0 30000 60000 90000 120000 150000
Appl. Sci. 2019, 9, x FOR PEER REVIEW 9 of 18
Frequency [Hz]
0.0025
(b) 5770 Hz R1=1MΩ
R2=5MΩ
R3=10MΩ
R4=50MΩ
0.0020 R5=100MΩ
R6=500MΩ
Open-circuit
|Voltage FRFs| [V/g]
0.0015
0.0010
0.0005
0.0000
4000 4500 5000 5500 6000 6500 7000 7500
Frequency [Hz]
Figure 2.
Figure Voltagefrequency
2. Voltage frequencyresponse
responsefunctions
functions(FRFs)
(FRFs)ofofthe
theflexoelectric
flexoelectricbeam
beamwith
with33 μm thickness
µm thickness
over the frequency range of (a) the first three modes; (b) around the first
over the frequency range of (a) the first three modes; (b) around the first mode.mode.
When the flexoelectric beam was shrunk proportionally to 0.3 μm in thickness, the voltage
FRFs are plotted in Figure 3. It is clear to observe that the resonance frequency increases about 10
times compared to Figure 2 for the beam with thickness 3 μm, as expected by Equation (25). The
enlarged view of the voltage FRFs near the first mode is shown in Figure 3b. It is interesting to see
that the resonance frequency corresponding to the peak voltage FRFs shifts from the short circuit
resonance frequency (𝑓 = 57,703 Hz) to the open circuit resonance frequency (𝑓 = 61,809 Hz) as
4000 4500 5000 5500 6000 6500 7000 7500
Frequency [Hz]
Figure 2. Voltage frequency response functions (FRFs) of the flexoelectric beam with 3 μm thickness
Appl. over the frequency
Sci. 2019, 9, 2273 range of (a) the first three modes; (b) around the first mode. 10 of 18
When the flexoelectric beam was shrunk proportionally to 0.3 μm in thickness, the voltage
FRFs When the flexoelectric
are plotted in Figure 3. beam
It is was
clearshrunk proportionally
to observe to 0.3 µm frequency
that the resonance in thickness, the voltage
increases aboutFRFs
10
times compared to Figure 2 for the beam with thickness 3 μm, as expected by Equation (25). The
are plotted in Figure 3. It is clear to observe that the resonance frequency increases about 10 times
compared
enlarged to Figure
view of the2voltage
for the FRFs
beamnear
withthethickness 3 µm,
first mode is as expected
shown by Equation
in Figure 3b. It is(25). The enlarged
interesting to see
viewthe
that of resonance
the voltagefrequency
FRFs nearcorresponding
the first modetois the
shownpeakinvoltage
Figure FRFs
3b. It shifts
is interesting
from thetoshort
see that the
circuit
resonance frequency (𝑓 = 57,703 Hz) to the open circuit resonance frequency (𝑓 = 61,809 Hz) as
resonance frequency corresponding to the peak voltage FRFs shifts from the short circuit resonance
sc = 57, 703 Hz) to the open circuit resonance frequency ( f oc = 61, 809 Hz) as the loading
loading( fresistance
frequency
the 1 increases. The short circuit resonance frequencies 1 can be obtained by solving
Equations (23) and (25), and the open circuit resonance frequencies can by
resistance increases. The short circuit resonance frequencies can be obtained be solving
predictedEquations (23)
using the
and (25), result
previous and the open
[53]. circuit
This resultresonance
indicates frequencies can be predicted
that the resonance frequencyusing the previous
of a small-scale result [53].
flexoelectric
This result
energy indicatesdevice
harvesting that theisresonance
dependent frequency
on the ofexternal
a small-scale flexoelectric
loading resistance.energy harvesting
Indeed, device
a resonance
is dependent on the external loading resistance. Indeed, a resonance frequency
frequency shift was observed previously in piezoelectric and flexoelectric energy harvesting devices shift was observed
previously in piezoelectric and flexoelectric energy harvesting devices [37,41,42,54].
[37,41,42,54].
-2
10
(a) R1=1MΩ
R2=5MΩ
R3=10MΩ
-4
R4=50MΩ
10 R5=100MΩ
R6=500MΩ
Open-circuit
|Voltage FRFs| [V/g]
-6
10
-8
10
-10
10
-12
10
0 300000 600000 900000 1200000 1500000
Appl. Sci. 2019, 9, x FOR PEER REVIEW 10 of 18
Frequency [Hz]
0.00018
(b) R1=1MΩ
61809 Hz R2=5MΩ
R3=10MΩ
0.00015 R4=50MΩ
R5=100MΩ
R6=500MΩ
Open-circuit
0.00012
|Voltage FRFs| [V/g]
0.00009
0.00006
0.00003
57703 Hz
0.00000
50000 55000 60000 65000 70000
Frequency [Hz]
Figure 3. Voltage FRFs of the flexoelectric beam with 0.3 µm thickness over the frequency range of
Figure 3. Voltage FRFs of the flexoelectric beam with 0.3 μm thickness over the frequency range of
(a) the first three modes; (b) around the first mode.
(a) the first three modes; (b) around the first mode.
Figure 4 gives the variation of the voltage FRFs for the external excitations at the short circuit
Figure
resonance 4 gives and
frequency the variation of theresonance
the open circuit voltage FRFs for therespectively.
frequency, external excitations
In log-logat the both
scale, shortvoltage
circuit
resonance frequency and the open circuit resonance frequency, respectively. In log-log scale, both
voltage outputs increase with increasing loading resistance, and trend to constant values at large
loading resistance. The two curves intersect at a point where the loading resistance is about 97 MΩ.
If the loading resistance was greater than 97 MΩ, the voltage output at the open circuit resonance
frequency was larger because the system was closer to the open circuit condition. On the other hand,
for lower loading resistance, the voltage output at the short circuit resonance frequency was larger.
Frequency [Hz]
Figure 3. Voltage FRFs of the flexoelectric beam with 0.3 μm thickness over the frequency range of
(a) the first three modes; (b) around the first mode.
Appl. Sci. 2019, 9, 2273 11 of 18
Figure 4 gives the variation of the voltage FRFs for the external excitations at the short circuit
resonance frequency and the open circuit resonance frequency, respectively. In log-log scale, both
outputsoutputs
voltage increaseincrease
with increasing loadingloading
with increasing resistance, and trend
resistance, andtotrend
constant values at
to constant largeatloading
values large
resistance. The two curves intersect at a point where the loading resistance is
loading resistance. The two curves intersect at a point where the loading resistance is about about 97 MΩ. 97 IfMΩ.
the
Ifloading resistance
the loading was greater
resistance than 97than
was greater MΩ,97 theMΩ,
voltage output at
the voltage the open
output circuit
at the openresonance frequency
circuit resonance
was larger because the system was closer to the open circuit condition. On the
frequency was larger because the system was closer to the open circuit condition. On the otherother hand, forhand,
lower
loading resistance, the voltage output at the short circuit resonance frequency was larger.
for lower loading resistance, the voltage output at the short circuit resonance frequency was larger. Therefore,
when the loading
Therefore, when the resistance was largerwas
loading resistance thanlarger
this special value,
than this choosing
special value, the open circuit
choosing the openresonance
circuit
resonance frequency as the external excitation frequency would be optimal for the voltage FRFssystem
frequency as the external excitation frequency would be optimal for the voltage FRFs since the since
wassystem
the closer was
to the opentocircuit
closer condition
the open circuitand vice versa.
condition and vice versa.
-2
10
sc
f1 = 57703 Hz
oc
f1 = 61809 Hz
-4
10
|Voltage FRFs| [V/g]
-6
10
-8
10
-10
10
0.01 0.1 1 10 100 1000
Resistance [MΩ]
Figure 4. Variation of the voltage output with load resistance under excitations at the short and open
Figure 4. Variation of the voltage output with load resistance under excitations at the short and open
circuit resonance frequencies of the first vibration mode (h = 0.3 µm).
circuit resonance frequencies of the first vibration mode (ℎ = 0.3 μm).
4.2. Tip Displacement FRFs
4.2. Tip Displacement FRFs
The multi-modes expression for the tip displacement FRFs can be obtained as
jωKr λr
P3
3
w(L, t) X r=1 (ω2 −ω2 )+2jωξr ωr
∅r ( L )
r
λr −
= Kr . (36)
−ω2 W0 ejωt jωKr2
κ33 BL
P3 ω2 − ω2 + 2jωξ ω
1
R + h jω +
r=1 r
r r
r=1 ( ω2r −ω2)+2jωξr ωr
The tip displacement FRFs of the beam with the excitation frequency for the 3 µm thickness beam
with the tip mass Mt are plotted in Figure 5. For the different load resistances, the values of the tip
displacement FRFs were almost equal. In other words, the tip displacement FRFs were independent
of the loading resistance. It is interesting to see that in Figure 5, there are two small anti-resonance
frequencies between mode 2 and mode 3, which is different from the result of piezoelectric energy
harvesting devices [42]. In the piezoelectric case without the tip mass, there was just one intense
mechanical anti-resonance frequency between mode 2 and mode 3.
Figure 6 plots the tip displacement FRFs of the 0.3 µm thickness beam with the tip mass Mt
near the first mode. Comparing Figures 5 and 6, it is clear to observe that the tip displacement
FRFs depend on the external loading resistance for the small scale flexoelectric energy harvesting.
The significant resonance frequency shift from f1sc to f1oc could be observed with increasing load
resistance. Apparently, at a small scale, the electromechanical coupling of the flexoelectric energy
harvesting is enhanced, which was not the case for the piezoelectric energy harvesting. It is interesting
to see that the peak amplitude of the tip displacement FRFs first decreases and then increases with
∑
( , ) ∅ ( )
=∑ 𝜆 − 𝐾 . (36)
∑
The tip displacement FRFs of the beam with the excitation frequency for the 3 μm thickness beam
with the tip mass 𝑀 are plotted in Figure 5. For the different load resistances, the values of the tip
Appl. Sci. 2019, 9, 2273 12 of 18
displacement FRFs were almost equal. In other words, the tip displacement FRFs were independent
of the loading resistance. It is interesting to see that in Figure 5, there are two small anti-resonance
increasing load
frequencies resistance.
between mode 2This andwas due3,to
mode the power
which dissipation
is different as aresult
from the resultofofpiezoelectric
Joule heatingenergy
in the
electrical domain
harvesting deviceswith
[42].the
Infinite values of loadcase
the piezoelectric resistance
without[42].
theThis phenomenon
tip mass, there was canjust
alsoone
be found
intense in
the case of piezoelectric
mechanical anti-resonance energy harvesting
frequency systems.
between mode 2 and mode 3.
0
10 (a) Short-circuit
R1=1MΩ
R2=5MΩ
-1
10 R3=10MΩ
|Tip displacement FRFs| [μm/g]
R4=50MΩ
R5=100MΩ
R6=500MΩ
-2
10 Open-circuit
-3
10
-4
10
-5
10
-6
10
0 30000 60000 90000 120000 150000
Frequency [Hz]
Appl. Sci. 2019, 9, x FOR PEER REVIEW 12 of 18
Tipdisplacement
Figure5.5.Tip
Figure displacementFRFs the33 µm
FRFsofofthe μm thickness
thickness beam
beam for
for the
the first
first three
three modes.
modes.
R5=100MΩ
The significant resonance frequency shift from 𝑓 to 𝑓 could be observed R6=500MΩ with increasing load
Open-circuit
resistance. Apparently, 0.0012
at a small scale, the electromechanical coupling of the flexoelectric energy
harvesting is enhanced, which was not the case for the piezoelectric energy harvesting. It is
interesting to see that the peak amplitude of the tip displacement FRFs first decreases and then
0.0009
increases with increasing load resistance. This was due to the power dissipation as a result of Joule
heating in the electrical 0.0006
domain with the finite values of load resistance [42]. This phenomenon can
also be found in the case of piezoelectric energy harvesting systems.
0.0003
0.0000
50000 55000 60000 65000 70000
Frequency [Hz]
Figure 6. Tip displacement FRFs of the 0.3 µm thickness beam with the tip mass Mt for the first mode.
Figure 6. Tip displacement FRFs of the 0.3 μm thickness beam with the tip mass 𝑀 for the first
mode.Density FRFs
4.3. Power
For energy
4.3. Power harvesting,
Density FRFs the power density is an important measure of performance. In the case of
the flexoelectric beam, the power density FRFs could be expressed as follows
For energy harvesting, the power density is an important measure of performance. In the case
of the flexoelectric beam, the power densityPFRFs could be expressed asfollows
3 −jωKr λr
2
(
Pd t ) r = 1 ( ω2 −ω2 +2jωξ ω
) 1 1
r r r
= , (37)
∑ jωKr2
−ω2 W0 ejωt ( ) 1 + κ33 BL jω + P3
R Vol
=
R h r=1 (ω2 −ω2 )+2jωξr ωr
, (37)
∑ r
where 𝑉𝑜𝑙 = 𝐵 × 𝐿 × ℎ is the volume of the beam. Figure 7 shows the power density FRFs of the
flexoelectric beams for the first mode. The frequency of peak power density is almost unchanged in
Figure 7a for the 3 μm thickness beam. The optimum power density was achieved with a finite load
resistance of ~100 MΩ. For the 0.3 μm thickness beam, Figure 7b shows the variation of the power
density FRFs with the various load resistances for the external excitations. Due to enhanced
electromechanical coupling in the small scale, the resonance frequency of the power density shifts
Appl. Sci. 2019, 9, 2273 13 of 18
where Vol = B × L × h is the volume of the beam. Figure 7 shows the power density FRFs of the
flexoelectric beams for the first mode. The frequency of peak power density is almost unchanged
in Figure 7a for the 3 µm thickness beam. The optimum power density was achieved with a finite
load resistance of ∼ 100 MΩ. For the 0.3 µm thickness beam, Figure 7b shows the variation of the
power density FRFs with the various load resistances for the external excitations. Due to enhanced
electromechanical coupling in the small scale, the resonance frequency of the power density shifts
from f1sc to f1oc with increasing load resistance, which is consistent with the result of Figure 3b.
The corresponding peak power density increases to a certain value (at frequency ~61,084 Hz and the
load resistance ~200 MΩ) and then decreases with increasing load resistance. Comparing the results of
Figure 7a,b, it should be noted that the maximum power density for the 0.3 µm thickness beam was
around 2 times that of the 3 µm thickness beam. This suggests that the flexoelectric energy harvesting
was enhanced in the smaller scale. Compared to piezoelectric energy harvesting, the peak power
density of the flexoelectric energy harvesting was dependent on the structure size. Hence, we could
find the optimal structure parameters to obtain the maximum power density.
Appl. Sci. 2019, 9, x FOR PEER REVIEW 13 of 18
1.0
(a) R1=1MΩ
R2=5MΩ
5770 Hz
R3=10MΩ
R4=50MΩ
0.8 R5=100MΩ
|Power density FRFs| [(W/m )/g ]
2
R6=500MΩ
3
Open-circuit
0.6
0.4
0.2
0.0
4000 4500 5000 5500 6000 6500 7000 7500
Frequency [Hz]
1.6
(b) R1=1MΩ
1.2 R 6=200MΩ
3
R7=500MΩ
Open-circuit
1.0
0.8
57703 Hz
0.6
0.4
0.2
0.0
50000 55000 60000 65000 70000
Frequency [Hz]
Powerdensity
7. Power
Figure 7. densityFRFs
FRFsofof
thethe beam
beam forfor
thethe
firstfirst order
order mode
mode (a) ℎ
for for (a)=h3 =μm
3 µm thickness;
thickness; (b)
ℎ = 0.3 μm thickness beams.
(b) h = 0.3 µm thickness beams.
The behaviors of power density FRFs with various load resistances for the external excitations
at the short and open circuit resonance frequencies are shown in Figure 8a. In log-log scale, the power
density increases linearly with increasing load resistance for low values of load resistance, and then
decreases for high values of load resistance. For low load resistances, the power density at the short
circuit resonance frequency was larger than that at the open circuit resonance frequency. For high
Appl. Sci. 2019, 9, 2273 14 of 18
The behaviors of power density FRFs with various load resistances for the external excitations at
the short and open circuit resonance frequencies are shown in Figure 8a. In log-log scale, the power
density increases linearly with increasing load resistance for low values of load resistance, and then
decreases for high values of load resistance. For low load resistances, the power density at the short
circuit resonance frequency was larger than that at the open circuit resonance frequency. For high
values of load resistance, there was the opposite phenomenon. There was an intersect point at the
load resistance of about 100 MΩ. The solid symbols in Figure 8a indicate the peak power density with
various resonance frequencies corresponding to the load resistances as shown in Figure 7b. For low
values of load resistance, the peak values were almost equal to those at the short circuit resonance
frequency. For high values of load resistance, the peak power density values were almost equal to
those at the open circuit resonance frequency. However, in the transition region of about 100 MΩ,
the optimal power density at the resonance frequency was larger than those at both the short and open
circuit resonance frequencies. Figure 8b plots the variation of the optimum external frequencies with
the load resistances. When the load resistances locate between 10 MΩ and 1000 MΩ, the optimum
external frequencies should be greater than the short circuit resonance frequency and smaller than the
open circuit resonance frequency.
Appl. Sci. 2019, 9, x FOR PEER REVIEW 14 of 18
1
10
(a)
0
10
|Power density FRFs| [(W/m )/g ]
2
3
-1
10
-2
10
sc
f1 = 57703 Hz
-3
10
oc
f1 = 61809 Hz
real frequency
-4
10
0.1 1 10 100 1000
Resistance [MΩ]
Figure 8. (a)8.Variation
Figure of the
(a) Variation power
of the powerdensity
densityFRFs
FRFs with loadresistance
with load resistanceforfor excitations
excitations at short
at the the short
and and
openopen
circuit resonance
circuit frequencies
resonance frequenciesofofthe
thefirst
first order mode(ℎ(h==0.30.3μm).
order mode µm). (b)(b) Variation
Variation of optimal
of the the optimal
external frequencies
external with
frequencies thethe
with load resistances.
load resistances.
Figure 9a gives the variation of the peak power density FRFs with beam thickness for the
different flexoelectric coefficients. With decreasing thickness, the peak power density FRFs increases
first and then decreases. When the thickness was small, the relatively large induced potential due to
the large strain gradient opposes the mechanical bending of the beam, which would reduce the beam
bending. When the thickness was large, the induced potential could not increase sufficiently to
Appl. Sci. 2019, 9, x FOR PEER REVIEW 15 of 18
Appl. Sci. 2019, 9, 2273 15 of 18
the maximum power density in the flexoelectric energy harvesting was about four times as long as
ℎ for the induced electric potential in the flexoelectric sensor. Figure 9b shows the optimal resistance
Figure 9a gives the variation of the peak power density FRFs with beam thickness for the different
for the peak power density as a function of the beam thickness for different flexoelectric coefficients.
flexoelectric coefficients. With decreasing thickness, the peak power density FRFs increases first and
When the thickness was large, the optimal resistance of the flexoelectric energy harvesting was nearly
then decreases. When the thickness was small, the relatively large induced potential due to the large
independent of the thickness. It is interesting that the optimal resistance (about 110 MΩ )
strain gradient opposes the mechanical bending of the beam, which would reduce the beam bending.
corresponding to the large thickness was almost independent on the flexoelectric coefficient. For a
When the thickness was large, the induced potential could not increase sufficiently to overcome the
particular flexoelectric coefficient, there was a critical thickness, below which the optimal resistance
increasing volume due to small strain gradient in elastic deformation. For a particular flexoelectric
increases with decreasing thickness. The critical thickness decreases with decreasing flexoelectric
coefficient, there was an optimal size for the flexoelectric energy harvesting, which could achieve
coefficient. Therefore, Figure 9 could provide optimal design parameters in terms of beam thickness
the maximum power density. This optimal size increases with increasing flexoelectric coefficient.
and load resistance for flexoelectric energy harvesting. It should be noted that, although the small
Our previous work [33] obtained the thickness h0 for the maximum induced electric potential in the
structure considered in the present study may not be applicable for the ambient vibration energy
flexoelectric sensors:
harvesting because of its super high frequency,v
t it is possible to design large size PVDF energy
12µ23113
harvesting with giant flexoelectricity [55]
h0 =for the ambient vibration
. frequency because the optimal
(38)
c1111 κ33 + e2311
thickness for the maximum power density is proportional to the flexoelectric coefficient.
Figure 9. (a)
Figure 9. (a) Variation
Variation of
of the
the peak
peak power
power density
density FRFs
FRFs and
and (b)
(b) variation
variation of
of the
the corresponding
corresponding optimal
optimal
resistance with beam thicknesses at different flexoelectric coefficients.
resistance with beam thicknesses at different flexoelectric coefficients.
Appl. Sci. 2019, 9, 2273 16 of 18
When flexoelectric coefficients were 1.15 × 10−2 , 2.3 × 10−2 , and 4.6 × 10−2 µC/m, the thicknesses
h0 were about 0.07, 0.14, and 0.28 µm, respectively. Figure 9a shows that the optimal thickness for the
maximum power density in the flexoelectric energy harvesting was about four times as long as h0
for the induced electric potential in the flexoelectric sensor. Figure 9b shows the optimal resistance
for the peak power density as a function of the beam thickness for different flexoelectric coefficients.
When the thickness was large, the optimal resistance of the flexoelectric energy harvesting was
nearly independent of the thickness. It is interesting that the optimal resistance (about 110 MΩ)
corresponding to the large thickness was almost independent on the flexoelectric coefficient. For a
particular flexoelectric coefficient, there was a critical thickness, below which the optimal resistance
increases with decreasing thickness. The critical thickness decreases with decreasing flexoelectric
coefficient. Therefore, Figure 9 could provide optimal design parameters in terms of beam thickness and
load resistance for flexoelectric energy harvesting. It should be noted that, although the small structure
considered in the present study may not be applicable for the ambient vibration energy harvesting
because of its super high frequency, it is possible to design large size PVDF energy harvesting with
giant flexoelectricity [55] for the ambient vibration frequency because the optimal thickness for the
maximum power density is proportional to the flexoelectric coefficient.
5. Conclusions
The flexoelectric energy harvesting based on the Euler–Bernoulli beam model has been investigated
in this paper. Using the electric Gibbs free energy and the generalized Hamilton’s principle for the
flexoelectric body, the electromechanically-coupled dynamics and electrical circuit equations of the
flexoelectric cantilever beams have been developed. The mode-superposition method was used to
obtain the closed-form analytical expressions of the electrical and mechanical responses in the modal
space. The numerical results indicate that the frequency of the peak vibration response exhibits a shift
from low frequency to high frequency with increasing load resistance in the small scale. The peak
voltage and power density with finite values of the loading resistance depend on the external excitations.
Interestingly, there is an intrinsic length scale for the optimal power density, which depends on the
material properties. Different from piezoelectricity, flexoelectric energy harvesting has an enhancement
of the electromechanical coupling with decreasing size and thus could be more effective in micro- or
nanoscale electromechanical systems.
Author Contributions: Conceptualization, R.H. and Z.Z.; Methodology, Y.S., X.L., and Z.Z.; Investigation and
data analysis, Y.S., X.L., and Z.Z.; Writing—original-draft, Y.S. and X.L.; Writing—review and editing, Y.S., R.H.,
and Z.Z.
Funding: This research was funded by the National Natural Science Foundation of China (Grant No. 11572271),
Young Teachers Education and Research Projects of Fujian Province (Grant No. JAT170918), Youth Foundation
of Chengyi University College, Jimei University (No. CK17002), and Scientific and Technological Innovation
Platform of Fujian Province (2006L2003).
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Hudak, N.S.; Amatucci, G.G. Small-scale energy harvesting through thermoelectric, vibration, and
radiofrequency power conversion. J. Appl. Phys. 2008, 103, 101301. [CrossRef]
2. Williams, C.B.; Yates, R.B. Analysis of a micro-electric generator for microsystems. Sens. Actuat. A Phys.
1996, 52, 8–11. [CrossRef]
3. Glynne-Jones, P.; Tudor, M.J.; Beeby, S.P.; White, N.M. An electromagnetic, vibration-powered generator for
intelligent sensor systems. Sens. Actuat. A Phys. 2004, 110, 344–349. [CrossRef]
4. Mathúna, C.O.; O’Donnell, T.; Martinez-Catala, R.V.; Rohan, J.; O’Flynn, B. Energy scavenging for long-term
deployable wireless sensor networks. Talanta 2008, 75, 613–623. [CrossRef] [PubMed]
5. Mitcheson, P.D.; Miao, P.; Stark, B.H.; Yeatman, E.M.; Holmes, A.S.; Green, T.C. MEMS electrostatic
micropower generator for low frequency operation. Sens. Actuat. A Phys. 2004, 115, 523–529. [CrossRef]
Appl. Sci. 2019, 9, 2273 17 of 18
6. Roundy, S.; Wright, P.K.; Rabaey, J. A study of low level vibrations as a power source for wireless sensor
nodes. Comput. Commun. 2003, 26, 1131–1144. [CrossRef]
7. Jeon, Y.B.; Sood, R.; Jeong, J.H.; Kim, S.G. MEMS power generator with transverse mode thin film PZT.
Sens. Actuat. A Phys. 2005, 122, 16–22. [CrossRef]
8. Anton, S.R.; Sodano, H.A. A review of power harvesting using piezoelectric materials (2003–2006). Smart Mater.
Struct. 2007, 16, R1–R21. [CrossRef]
9. Lu, C.; Tsui, C.Y.; Ki, W.H. A batteryless vibration-based energy harvesting system for ultra low power
ubiquitous applications. IEEE Int. Sym. Circuit Syst. 2007, 1349–1352.
10. Lu, C.; Tsui, C.Y.; Ki, W.H. Vibration energy scavenging system with maximum power tracking for micropower
applications. IEEE Trans. VLSI Syst. 2011, 19, 2109–2119. [CrossRef]
11. Lu, C.; Raghunathan, V.; Roy, K. Efficient design of micro-scale energy harvesting systems. IEEE J. Emerg.
Sel. Top. Circuit Syst. 2011, 1, 254–266. [CrossRef]
12. Ma, W.; Cross, L.E. Large flexoelectric polarization in ceramic lead magnesium niobate. Appl. Phys. Lett.
2001, 79, 4420–4422. [CrossRef]
13. Ma, W.; Cross, L.E. Strain-gradient-induced electric polarization in lead zirconate titanate ceramics.
Appl. Phys. Lett. 2003, 82, 3293–3295. [CrossRef]
14. Zubko, P.; Catalan, G.; Buckley, A.; Welche, P.R.L.; Scott, J.F. Strain-gradient-induced polarization in SrTiO3
single crystals. Phys. Rev. Lett. 2007, 99, 167601. [CrossRef]
15. Sharma, N.D.; Maranganti, R.; Sharma, P. On the possibility of piezoelectric nanocomposites without using
piezoelectric materials. J. Mech. Phys. Solids 2007, 55, 2328–2350. [CrossRef]
16. Sharma, N.D.; Landis, C.M.; Sharma, P. Piezoelectric thin-film superlattices without using piezoelectric
materials. J. Appl. Phys. 2010, 108, 024304. [CrossRef]
17. Majdoub, M.S.; Sharma, P.; Cagin, T. Enhanced size-dependent piezoelectricity and elasticity in nanostructures
due to the flexoelectric effect. Phys. Rev. B 2008, 77, 125424. [CrossRef]
18. Qi, Y.; Kim, J.; Nguyen, T.D.; Lisko, B.; Purohit, P.K.; McAlpine, M.C. Enhanced piezoelectricity and
stretchability in energy harvesting devices fabricated from buckled PZT ribbons. Nano Lett. 2011, 11, 1331.
[CrossRef]
19. Zhu, W.; Fu, J.Y.; Li, N.; Cross, L. Piezoelectric composite based on the enhanced flexoelectric effects.
Appl. Phys. Lett. 2006, 89, 192904. [CrossRef]
20. Chu, B.; Zhu, W.; Li, N.; Cross, L.E. Flexure mode flexoelectric piezoelectric composites. J. Appl. Phys. 2009,
106, 104109. [CrossRef]
21. Zubko, P.; Catalan, G.; Tagantsev, A.K. Flexoelectric effect in solids. Ann. Rev. Mater. Res. 2013, 43, 387–421.
[CrossRef]
22. Lee, D.; Noh, T.W. Giant flexoelectric effect through interfacial strain relaxation. Philos. Trans. R. Soc. A Math.
Phys. Eng. Sci. 2012, 370, 4944–4957. [CrossRef]
23. Petrov, A.G. Electricity and mechanics of biomembrane systems: Flexoelectricity in living membranes.
Anal. Chim. Acta 2006, 568, 70–83. [CrossRef]
24. Abdollahi, A.; Peco, C.; Millan, D.; Arroyo, M.; Arias, I. Computational evaluation of the flexoelectric effect
in dielectric solids. J. Appl. Phys. 2014, 116, 093502. [CrossRef]
25. Hu, S.L.; Sheng, S.P. Electric field gradient theory with surface effect for nano-dielectrics. Comput. Mater. Conit.
2009, 13, 63–87.
26. Zhang, Z.; Yan, Z.; Jiang, L. Flexoelectric effect on the electroelastic responses and vibrational behaviors of a
piezoelectric nanoplate. J. Appl. Phys. 2014, 116, 014307. [CrossRef]
27. He, L.; Lou, J.; Zhang, A.; Wu, H.; Du, J.; Wang, J. On the coupling effects of piezoelectricity and flexoelectricity
in piezoelectric nanostructures. AIP Adv. 2017, 7, 105106. [CrossRef]
28. Xiang, S.; Li, X.F. Elasticity solution of the bending of beams with the flexoelectric and piezoelectric effect.
Smart Mater. Struct. 2018, 27, 105023. [CrossRef]
29. Su, Y.X.; Zhou, Z.D.; Yang, F.P. Electromechanical analysis of bilayer piezoelectric sensors due to flexoelectricity
and strain gradient elasticity. AIP Adv. 2019, 9, 015207. [CrossRef]
30. Toupin, R.A. The elastic dielectric. Arch. Ration. Mech. Anal. 1956, 5, 849–915. [CrossRef]
31. Shen, S.P.; Hu, S.L. A theory of flexoelectricity with surface effect for elastic dielectrics. J. Mech. Phys. Solids
2010, 58, 655–677. [CrossRef]
Appl. Sci. 2019, 9, 2273 18 of 18
32. Liang, X.; Hu, S.; Shen, S. Effects of surface and flexoelectricity on a piezoelectric nanobeam. Smart Mater. Struct.
2014, 23, 035020. [CrossRef]
33. Yan, Z.; Jiang, L.Y. Flexoelectric effect on the electroelastic responses of bending piezoelectric nanobeams.
J. Appl. Phys. 2013, 113, 194102. [CrossRef]
34. Liang, X.; Hu, S.; Shen, S. Size-dependent buckling and vibration behaviors of piezoelectric nanostructures
due to flexoelectricity. Smart Mater. Struct. 2015, 24, 105012. [CrossRef]
35. Wang, X.; Zhang, R.; Jiang, L. A study of the flexoelectric effect on the electroelastic fields of a cantilevered
piezoelectric nanoplate. Int. J. Appl. Mech. 2017, 9, 1750056. [CrossRef]
36. Zhou, Z.D.; Yang, C.P.; Su, Y.X.; Huang, R.; Lin, X.H. Electromechanical coupling in piezoelectric nanobeams
due to flexoelectric effect. Smart Mater. Struct. 2017, 26, 095025. [CrossRef]
37. Deng, Q.; Kammoun, M.; Erturk, A.; Sharma, P. Nanoscale flexoelectric energy harvesting. Int. J. Solids Struct.
2014, 51, 3218–3225. [CrossRef]
38. Moura, A.G.; Erturk, A. Electroelastodynamics of flexoelectric energy conversion and harvesting in elastic
dielectrics. J. Appl. Phys. 2017, 121, 064110. [CrossRef]
39. Yan, Z. Modeling of a nanoscale flexoelectric energy harvester with surface effects. Physical E 2017, 88,
125–132. [CrossRef]
40. Liang, X.; Zhang, R.; Hu, S.; Shen, S. Flexoelectric energy harvesters based on Timoshenko laminated beam
theory. J. Intell. Mater. Syst. Strust. 2017, 28, 2064–2073. [CrossRef]
41. Erturk, A.; Inman, D.J. An experimentally validated bimorph cantilever model for piezoelectric energy
harvesting from base excitations. Smart Mater. Struct. 2009, 18, 025009. [CrossRef]
42. Erturk, A.; Inman, D.J. Piezoelectric Energy Harvesting; Wiley: Hoboken, NJ, USA, 2011.
43. Tang, L.; Wang, J. Size effect of tip mass on performance of cantilevered piezoelectric energy harvester with a
dynamic magnifier. Acta Mech. 2017, 228, 3997–4015. [CrossRef]
44. Chu, B.; Salem, D.R. Flexoelectricity in several thermoplastic and thermosetting polymers. Appl. Phys. Lett.
2012, 101, 103905. [CrossRef]
45. Daqaq, M.F.; Stabler, C.; Qaroush, Y.; Seuaciuc-Osorio, T. Investigation of power harvesting via parametric
excitations. J. Intell. Mater. Syst. Struct. 2009, 20, 545–557. [CrossRef]
46. Ferrari, M.; Ferrari, V.; Guizzetti, M.; Ando, B.; Baglio, S.; Trigona, C. Improved energy harvesting from wideband
vibrations by nonlinear piezoelectric converters. Sens. Actuat. A Phys. 2010, 162, 425–431. [CrossRef]
47. Mahmoudi, S.; Kacem, N.; Bouhaddi, N. Enhancement of the performance of a hybrid nonlinear vibration
energy harvester based on piezoelectric and electromagnetic transductions. Smart Mater. Struct. 2014,
23, 075024. [CrossRef]
48. Drezet, C.; Kacem, N.; Bouhaddi, N. Design of a nonlinear energy harvester based on high static low dynamic
stiffness for low frequency random vibrations. Sens. Actuat. A Phys. 2018, 283, 54–64. [CrossRef]
49. Yang, B.; Lee, C.; Xiang, W.; Xie, J.; He, J.H.; Kotlanka, R.K.; Low, S.P.; Feng, H. Electromagnetic energy
harvesting from vibrations of multiple frequencies. J. Micromech. Microeng. 2009, 19, 035001. [CrossRef]
50. Sari, I.; Balkan, T.; Kulah, H. An electromagnetic micro power generator for wideband environmental
vibrations. Sens. Actuat. A Phys. 2008, 145–146, 405–413. [CrossRef]
51. Abed, I.; Kacem, N.; Bouhaddi, N.; Bouazizi, M.L. Multi-modal vibration energy harvesting approach based
on nonlinear oscillator arrays under magnetic levitation. Smart Mater. Struct. 2016, 25, 025018. [CrossRef]
52. Abed, I.; Kacem, N.; Bouhaddi, N.; Bouazizi, M.L. Nonlinear dynamics of magnetically coupled beams for
multi-modal vibration energy harvesting. In Proceedings of the 2016 SPIE Smart Structures and Materials +
Nondestructive Evaluation and Health Monitoring, Las Vegas, NV, USA, 15 April 2016; p. 97992C.
53. Lin, X.H.; Su, Y.X.; Zhou, Z.D.; Yang, J.P. Analysis of the natural frequency for flexoelectric cantilever beams
under the open-circuit condition. Chin. Q. Mech. 2018, 39, 383–394. (In Chinese)
54. Dutoit, N.E.; Wardle, B.L. Experimental verification of models for microfabricated piezoelectric vibration
energy harvesters. AIAA J. 2007, 45, 1126–1137. [CrossRef]
55. Baskaran, S.; Ramachandran, N.; He, X.; Thiruvannamalai, S.; Lee, H.J.; Heo, H.; Chen, Q.; Fu, J.Y.
Giant flexoelectricity in polyvinylidene fluoride films. Phys. Lett. A 2011, 375, 2082–2084. [CrossRef]
© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).