You are on page 1of 19

Nanoscale piezoelectric vibration energy

harvester design
Cite as: AIP Advances 7, 095122 (2017); https://doi.org/10.1063/1.4994577
Submitted: 05 July 2017 . Accepted: 28 August 2017 . Published Online: 26 September 2017

Hamid Reza Foruzande, Ali Hajnayeb, and Amin Yaghootian

ARTICLES YOU MAY BE INTERESTED IN

Energy harvesting from low frequency applications using piezoelectric materials


Applied Physics Reviews 1, 041301 (2014); https://doi.org/10.1063/1.4900845

Piezoelectric energy harvesting: State-of-the-art and challenges


Applied Physics Reviews 1, 031104 (2014); https://doi.org/10.1063/1.4896166

A utility piezoelectric energy harvester with low frequency and high-output voltage:
Theoretical model, experimental verification and energy storage
AIP Advances 6, 095208 (2016); https://doi.org/10.1063/1.4962979

AIP Advances 7, 095122 (2017); https://doi.org/10.1063/1.4994577 7, 095122

© 2017 Author(s).
AIP ADVANCES 7, 095122 (2017)

Nanoscale piezoelectric vibration energy


harvester design
Hamid Reza Foruzande, Ali Hajnayeb,a and Amin Yaghootian
Department of Mechanical Engineering, Faculty of Engineering, Shahid Chamran
University of Ahvaz, Ahvaz 6135743337, Iran
(Received 5 July 2017; accepted 28 August 2017; published online 26 September 2017)

Development of new nanoscale devices has increased the demand for new types of
small-scale energy resources such as ambient vibrations energy harvesters. Among
the vibration energy harvesters, piezoelectric energy harvesters (PEHs) can be easily
miniaturized and fabricated in micro and nano scales. This change in the dimensions
of a PEH leads to a change in its governing equations of motion, and consequently,
the predicted harvested energy comparing to a macroscale PEH. In this research,
effects of small scale dimensions on the nonlinear vibration and harvested voltage of a
nanoscale PEH is studied. The PEH is modeled as a cantilever piezoelectric bimorph
nanobeam with a tip mass, using the Euler-Bernoulli beam theory in conjunction
with Hamilton’s principle. A harmonic base excitation is applied as a model of the
ambient vibrations. The nonlocal elasticity theory is used to consider the size effects
in the developed model. The derived equations of motion are discretized using the
assumed-modes method and solved using the method of multiple scales. Sensitivity
analysis for the effect of different parameters of the system in addition to size effects is
conducted. The results show the significance of nonlocal elasticity theory in the predic-
tion of system dynamic nonlinear behavior. It is also observed that neglecting the size
effects results in lower estimates of the PEH vibration amplitudes. The results pave
the way for designing new nanoscale sensors in addition to PEHs. © 2017 Author(s).
All article content, except where otherwise noted, is licensed under a Creative
Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).
https://doi.org/10.1063/1.4994577

I. INTRODUCTION
Several energy harvesters for powering nano-scale portable electronic systems have been devel-
oped as the demand for self-powered electronic devices grows rapidly.1,2 One of the most common
types of energy harvesters is piezoelectric harvesters. Piezoelectric energy harvesters have been
studied and applied in nanoscale extensively because of their simpler structures, scalability and
higher energy density in small scales.3,4 Wang4 reviewed the theoretical and experimental aspects
of previous research on piezoelectric nanowires. Chen et al.5 used piezoelectric nanofibers for
energy harvesting. Wang and Wang6 studied a nanoscale unimorph PEH. Wang and Wang7 and
Deng et al.8 studied the flexoelectric effects in nanoscale PEHs. After determining the piezoelec-
tric nano-generators as one of the best type of energy harvester in nanoscale, the next step is
modeling.
Expensive fabrication processes in nanoscale makes it more necessary to prepare an accurate
model during the designing process of a nanoscale device. Because of similarity between the structure
and elements of a PEH in nano and macro scales, the basis of modeling process in nanoscale is the
same as macro scale. Thereby, before modeling a nanoscale PEH, a review on the techniques of
modeling macroscale PEHs is necessary.

a
Corresponding author: P.O.B. 6135743337 Ahvaz, Iran, a.nayeb@scu.ac.ir

2158-3226/2017/7(9)/095122/18 7, 095122-1 © Author(s) 2017


095122-2 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

The modeling of macro scale PEHs has been investigated extensively in previous studies.9–13 A
typical PEH is comprised of a cantilevered piezoelectric beam with a tip mass, where the clamped
end of the beam is attached to the host vibrating structure. Erturk and Inman14 modeled a unimorph
PEH as a continuous system using Euler-Bernoulli beam theory, and compared its vibrations with
the simple single-degree-of-freedom (SDOF) models of previous research works. They showed that
the SDOF model may predict highly inaccurate vibrations comparing to the beam model. Although a
PEH with linear behavior is easily modeled and analyzed, but they have serious shortcomings such as
limited performance bandwidth. High values of output voltage in linear PEHs are limited to a narrow
band of frequency close to the resonance of the PEH. The simplest solution to low output voltage in
any vibration frequency is increasing the amplitude of vibrations. It does not only boost the output
voltage of a PEH in all frequencies, but also broadens its bandwidth as a result of nonlinear behavior.
Therefore, nonlinearity and size which is also an important issue in modeling, are discussed in the
following lines.
There are several studies on nonlinear vibrations of PEHs.15–17 Sebald et al.16 experimentally
studied the harvested power and bandwidth of a PEH with nonlinearities in stiffness and damping and
compared them with the performance of a linear PEH. The results showed significant improvements
in the bandwidth and harvested power. The type of stability also affects the performance of a PEH.
Cottone et al.15 theoretically and experimentally studied the nonlinear vibrations of a bistable oscil-
lator as a PEH. The PEH consisted of an axially loaded piezoelectric beam. They observed a superior
performance of the proposed PEH under a wideband random excitation. In many studies, an external
nonlinear force like the force of a permanent magnet is applied to induce a nonlinearity and use its
benefits in the PEH. Tang et al.17 used magnetic force to investigate influence of both monostable
and bistable configurations in a PEH under random excitations. They proved the superiority of these
configurations over linear PEHs. In summary, the nonlinearity boosts the performance of a PEH and
is desired in energy harvesting.
In addition to nonlinearity, another important factor in modeling of PEHs is the size of the struc-
ture. Molecular dynamics simulations and experiments show that the mechanical behaviors of micro
and nanoscale structures are significantly affected by size effects.18 Therefore, the classical continuum
theory as a scale-independent theory is unable to precisely model small structures. Recent studies sug-
gest using size-dependent theories (e.g. the modified couple stress theory, the strain gradient theory
and etc.).19 The nonlocal elasticity theory introduced by Eringen20 is one of the size-dependent theo-
ries that has been used successfully in the majority of previous research. The theory is also applicable
to piezoelectric materials and structures. It has been used to study the vibrations of the nanobeam in
a limited number of research works. Ke and Wang21 applied the nonlocal elasticity theory to study
the linear vibrations of a piezoelectric nanobeam. Differential quadrature (DQ) method was used to
obtain the natural frequencies and mode shapes of the nanobeam with different boundary conditions.
It was found that the applied continuum theory has an important role in finding more realistic values
for the natural frequencies. Ke et al.22 investigated the nonlinear vibrations of a nanobeam by using
nonlocal elasticity theory. The DQ method and a direct iterative method were used to obtain the non-
linear natural frequencies of the nanobeam. Because piezoelectric beams in small scales are mostly
fabricated in unimorph or bimorph forms, modeling bimorph and unimorph nanobeams in addition to
applying appropriate continuum theory is necessary. Nazemizadeh and Bakhtiari-nejad23 analytically
investigated the free vibrations of nano/microbeams using the nonlocal elasticity theory. The effects
of nonlocal parameter in addition to different parameters of the system on the natural frequencies
and mode shapes of the beam were studied. In conclusion, the resulting error from neglecting the
size effects in modeling a nanoscale PEH is significant.
To the best knowledge of authors no study has been published on the dynamic modeling of
nanoscale vibratory PEHs. They are used in powering small-scale electronic device and have to be
modeled before fabrication. Large amplitude nonlinear vibrations of a nanoscale PEH is investigated
in this paper. Nanoscale PEH is considered to be a piezoelectric bimorph cantilever nanobeam with a
tip mass. The nanobeam is subjected to harmonic base excitation. It is modeled using Euler-Bernoulli
beam theory combined with nonlocal elasticity theory to address the size effects in nanoscale. The
nonlinear equations of motion and related boundary conditions are derived employing Hamilton’s
principle. Then, the obtained equations are discretized using assumed modes method and solved
095122-3 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

utilizing the multiple scales method. To validate the obtained approximate solutions, they are com-
pared with a numeric solution, which shows good agreement. The results show the behavior of the
nanoscale PEH for different values of system parameters.

II. THEORETICAL MODELING


A bimorph piezoelectric nanobeam of length L and width b has been considered (figure 1). The
thickness of the piezoelectric layers is assumed to be hp while he indicates the thickness of the elastic
layer. Also, polarity direction of the piezoelectric layers is considered parallel to the z axis. In order to
consider direct and inverse effects of piezoelectric material concurrently, electric potential is assumed
as a summation of triogonometric and linear terms, according to Wang’s research:24
z1 V0
Φ(1) (x, z, t) = − cos ( βz1 ) φ (x, t) + exp(iΩt),
hp
(1)
z2 V0
Φ (x, z, t) = − cos ( βz2 ) φ (x, t) +
(3)
exp(iΩt),
hp

where Φ1 (x, z, t) and Φ3 (x, z, t) are total electric potential functions of the upper and lower piezo-
electric layers, respectively. The superscripts determines the layer number of the beam. Moreover, a
numerical value of β = hπp is known, and φ(x, t) indicates the generated voltage due to piezoelectric
direct effect. V0 is the amplitude of an external voltage and Ω is the frequency of this voltage. The
local coordinates of z1 and z2 are shown in figure 1. Because there is no external voltage in PEHs,
the second term in the right-hand-side of both Equations (1a) and (1b) is not considered in the rest
of this study. z1 and z2 can be written as the following functions of z:
1 
z1 = −z − he + hp ,
2
(2)
1 
z2 = −z + he + hp .
2
By substituting Equations (2) into Equations (1), the total electric potential functions can be
rearranged as:

FIG. 1. A schematic view of a) the studied PEH b) cross section.


095122-4 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)
" !#
1
Φ(1) (x, z, t) = − cos β −z − he + hp φ(x, t),
2
" (3)
1 !#
Φ(3) (x, z, t) = − cos β −z + he + hp φ(x, t).
2
After determining the total electric potential functions, the electric filed components can be obtained
as:
∂Φ(1)  ! # ∂φ
"
(1) 1
Ex = − = cos β −z − he + hp ,
∂x 2 ∂x

∂Φ (1)
∂Φ(1) ∂z
" !#
1
Ez(1) = − =− = − β sin β −z − he + hp φ,
∂z1 ∂z ∂z1 2
(4)
∂Φ (3)
∂φ
" !#
1
Ex(3) = − = cos β −z + he + hp ,
∂x 2 ∂x

∂Φ (3)
∂Φ(3) ∂z
" !#
1
Ez(3) = − =− = − β sin β −z + he + hp φ.
∂z2 ∂z ∂z2 2
Based on Euler-Bernoulli beam theory, the only nonzero component of the strain tensor of an element
of a beam is given as:
ε xx = ε ◦xx − zK, (5)
where ε ◦xx
is strain at neutral axis and K is beam curvature. ε ◦xx
is neglected because of applying
shortening assumption. Hence, the strain component of Equation (5) can be rewritten as:25
 
ε xx = −zK = −z W 00 − W 0U 00 − W 00U 0 − W 00W 02 , (6)

where U(x, t) and W (x, t) are the axial and transverse displacements of the nanobeam at point x and
time t, respectively.

III. NONLOCAL ELASTICITY THEORY


Based on Nonlocal elasticity theory, the constitutive equations of piezoelectric materials in
absence of body forces are expressed as:21

σij − (e0 a)2 ∇2 σij = Cijkl ε kl − ekij Ek ,


(7)
Di − (e0 a)2 ∇2 Di = eikl ε kl +  ik Ek ,
where σij , ε kl , Di and E k are stress, strain, electric displacement and electric filed components,
respectively. Also, C ijkl , ekij and  ik are elastic, piezoelectric and dielectric constants, respectively.
Since the thickness to the length ratio and the width to the length ration of the studied beam are
assumed to be small, Equation (7) can be rewritten in a reduced form as:

∂ 2 σxx
σxx − (e0 a)2 = C11 ε xx − e31 Ez , (8)
∂x 2

∂ 2 Dx
Dx − (e0 a)2 =  11 Ex , (9)
∂x 2

∂ 2 Dz
Dz − (e0 a)2 = e31 ε xx +  33 Ez . (10)
∂z2
095122-5 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

IV. DERIVING GOVERNING EQUATIONS


The strain energy of a bimorph piezoelectric nanobeam can be expressed as:
   he
1 b l − 2  (1)
σxx ε xx − Dx(1) Ex(1) − Dz(1) Ez(1) dzdx dy

Πs =
2 0 0 − 2e −hp
h

 b l he
+hp
1 2
(3)
ε xx − Dx(3) Ex(3) − Dz(3) Ez(3) dzdx dy
 
+ σxx (11)
2 0 0 he
2
 b l he
1 2  (2)

+ σxx ε xx dzdx dy.
2 0 0 − h2e

By defining the bending moments as below:


 − h2e  (1) 
M1 = b σxx z dz,
− h2e −hp
 he
2  (2) 
M2 = b σxx z dz, (12)
− h2e
 he
+hp
2  (3) 
M3 = b σxx z dz,
he
2

variation of the strain energy can be given as:


l
δΠs = (−M1 δk − M2 δk − M3 δk) dx
0
 b l − h2e
−Dx(1) δEx(1) + Dz(1) δEz(1) dzdx dy
 
+ (13)
0 0 − h2e −hp
 b l he
+hp
2
−Dx(3) δEx(3) + Dz(3) δEz(3) dzdx dy.
 
+
he
0 0 2

The kinetic energy of the nanobeam is defined as:



 ∂W 2
! 2
1 l ∂U 
! !
mtip
Πk = m1 + m2 + m3 + H(x − l2 )  + dx,
2 0 l2  ∂t ∂t 
 (14)
m1 = m3 = ρp bhp , m2 = ρe bhe ,
where ρp , ρe and mi are density of piezoelectric layers, density of elastic layer and mass per unit
length of the i-th layer of the nanobeam, respectively. Also, mtip and l 2 are the quantity and location
of the tip mass, respectively. In order to add a tip mass, Heaviside function dented by H(x) is used.
Therefore, the variation of kinetic energy is:

 ∂W 2
! 2
1 l ∂U 
!  !
mtip
δΠk = m1 + m2 + m3 + H(x − l2 δ 
) + dx. (15)
2 0 l2  ∂t ∂t 

The work done by external forces i.e. viscous damping force is expressed as:
l
∂W
Πf = C Wdx, (16)
0 ∂t
Where C is the viscous damping coefficient.The work done by the external forces is given as:
l
∂W
δΠf = C δWdx. (17)
0 ∂t
To derive equations of motion and related boundary conditions, Hamilton’s principle is employed.
Hamilton’s principle can be expressed as:
095122-6 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)
 t  t " l #
1 
0 2 02
δH = (δΠs + δΠf − δΠk )dt − λδ 1 − 1 + U −W dx = 0. (18)
0 2 0 0
The last term in Equation (18) is used to satisfy the inextensionality condition of the beam. λ is a
Lagrange multiplier and would be calculated later. By substituting Equations (13), (15) and (16) in
Equation (18) and then equating coefficients of δU, δW and δφ by zero, the equations of motion are
obtained:
λ 1 + U 0 0 = M1 W 0 00 + M2 W 0 00 + M3 W 0 00 − M1 W 00 0 − M2 W 00 0 − M3 W 00 0
      

mtip
! (19)
+ m1 + m2 + m3 + H(x − l2 ) Ü,
l2
−M1 00 − M2 00 − M3 00 − M1 U 00 0 − M2 U 00 0 − M3 U 00 0 + M1 U 0 00 + M2 U 0 00 + M3 U 0 00
     

   00
− 2 W 00W 0M1 0 − 2 W 00W 0M2 0 − 2 W 00W 0M3 0 + M1 W 02
 

!
  00   00 mtip (20)
+ M 2 W 0 2 + M 3 W 0 2 + m1 + m2 + m3 + H(x − l2 ) Ẅ
l2
∂W
− λW 0 0 = 0,

+C
∂t
 − h2e "  ! # ∂Dx(1) "  ! # 
Dz(1) β sin β −z − 1 he + hp 1
 
b + cos β −z − he + hp  dz
− h2e −hp  2 ∂x 2 
 he +hp " " !#
2
(3) 1
+b Dz β sin β −z + he + hp (21)
he
2
2

∂Dx(3)
"  ! # 
1
+ cos β −z + he + hp  dz = 0.
∂x 2 
Also boundary conditions at x = 0 and x = l are given as:
(M1 + M2 + M3 ) W 00 − (M1 + M2 + M3 ) W 0 0 + λ 1 + U 0 = 0, or U = 0,
  

(M1 + M2 + M3 ) W 0 = 0, or U = 0
 
M10 + M20 + M30 + (M1 + M2 + M3 ) U 00 − (M1 + M2 + M3 ) U 0 0
 

f g0
+ 2 (M1 + M2 + M3 ) W 00W 0 − (M1 + M2 + M3 ) W 02 + λW 0 = 0,

or W = 0,
− (M1 + M2 + M3 ) + (M1 + M2 + M3 ) U 0 + (M1 + M2 + M3 ) W 02 = 0, or W 0 = 0,
 he +hp " !#  − he " !#
2
(3) 1 2
(1) 1
Dx cos β −z + he + hp dz + Dx cos β −z − he + hp dz,
he
2
2 − h2e −hp 2
or φ = 0. (22)
It can be proved that the strain at neutral axis of the nanobeam is obtained as:25
q
ε ◦xx = (1 + U 0)2 + W 02 − 1. (23)
Using Taylor expansion, Equation (23) can be approximated as:
1
ε ◦xx = U 0 + W 02 . (24)
2
Shortening effect reduces the number of independent motion parameters from three to two by relating
the axial displacement to the transverse displacement as the following equation shows:26
1
ε ◦xx = 0 ⇒ U 0 = − W 02 , (25)
2
095122-7 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

and also:

1 ∂2 x !
U 00 = −W 0W 00, Ü = − W 02 dx . (26)
2 ∂t 2 0
λ is calculated by introducing Equations (25) and (26) in Equation (19):
"
1
λ= ! M1 w 0 0 + M2 w 0 0 + M3 w 0 0 − M1 w 00 − M2 w 00 − M3 w 00
  
1
1 − w 02
2 (27)
 ! 2 x
1 x ∂
! #
mtip 02
− m1 + m2 + m3 + H(x − l2 ) w dx dxdx .
2 l l2 ∂t 2 0
The denominator of Equation (27) can be substituted by its Taylor expansion as below:
!
1 1 1
! ≈ 1 + W 02 + W 04 . (28)
1 2 4
1 − W 02
2
Because of keeping the nonlinear terms up to the third order in this study, only the first term of the
expansion is kept and the other two terms are neglected. By substituting Equation (28) into Equation
(27), the Lagrange multiplier can be obtained as:
λ = M 0 1 + M 02 + M 03 w 0


 ! 2 x (29)
1 x ∂
!
mtip 02
− m1 + m2 + m3 + H(x − l2 ) w dx dxdx.
2 l l2 ∂t 2 0
By submitting Equations (25), (26) and (29) in Equation (20), Equation (30) is obtained:
1
−(M 001 + M 002 + M 003 ) − M 001 + M 002 + M 003 W 02 − M 01 + M 02 + M 03 W 0W 00 + C Ẇ
 
2
!
mtip
+ m1 + m2 + m3 + H(x − l2 ) Ẅ (30)
l2
" x ! x  #0
mtip 2

+ W 0
m1 + m2 + m3 + H(x − l2 ) Ẇ + W Ẅ dxdx = 0.
0 0 0
l l2 0

In order to achieve an explicit expression for the summation of second derivatives of the bending
moments, M 001 + M 002 + M 003 , higher order terms of Equation (30) are ignored:
M 001 + M 002 + M 003
!
mtip
= C Ẇ + m1 + m2 + m3 + H(x − l2 ) Ẅ
l2 (31)
" x ! x  #0
0 mtip 0 2 0 0

+ W m1 + m2 + m3 + H(x − l2 ) Ẇ + W Ẅ dxdx .
l l2 0

On other hand, by multiplying Equation (8) by z and then integrating on the thickness of each layer
one obtains the following equations:
!
(1) (1) 1
M1 − (e0 a)2 M 001 = F31 φ − D11 W 00 + W 00W 02 , (32-a)
2
!
(2) 1
M2 − (e0 a)2 M 002 = −D11 W 00 + W 00W 02 , (32-b)
2
!
(3) (3) 1
M3 − (e0 a)2 M 003 = F31 φ − D11 W 00 + W 00W 02 , (32-c)
2
095122-8 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

where
 − h2e !# "
(1) 1
F31 = be31 βzsin β −z − he + hp dz,
− h2e −hp 2
(33)
 he
+hp " !#
(3) 2 1
F31 = be31 βzsin β −z + he + hp dz,
he
2
2
 − h2e
(1) p
D11 = b C11 z2 dz,
− h2e −hp

 he
2
(2)
D11 = e 2
b C11 z dz, (34)
− h2e

 he
+hp
2
(3) p
D11 = b C11 z2 dz.
he
2

From Equation (9), the following equations can be obtained:


 − h2e 2 (1)
*Dx(1) − (e0 a)2 ∂ Dx + cos β −z − 1 he + hp (1) ∂φ
"  !#
b = X11 ,
− h2e −hp , ∂x -
2 2 ∂x
(35)
 he +hp 2 (3)
*Dx(3) − (e0 a)2 ∂ Dx + cos β −z + 1 he + hp (3) ∂φ
" !#
2 
b = X11 ,
he
2 , ∂x -
2 2 ∂x

where
 − h2e ( "  ! # )2
(1) 1
X11 = b ∈11 cos β −z − he + hp dz,
− h2e −hp 2
(36)
 he
+hp ( "  ! # )2
(3) 2 1
X11 = b ∈11 cos β −z + he + hp dz.
he
2
2
Similarly, following equations can be obtained using Equation (10):
 − h2e 2 (1) 
Dz(1) − (e0 a)2 ∂ Dz  βsin β −z − 1 he + hp (1) ∂ W
2
 " !# (1)

b = −F31 − X33 φ,
he
− 2 −hp   ∂z 2 

2 ∂x 2

(37)
 he +hp  2 (3) 
2 ∂ Dz  (3) ∂ W
2
" !#
2  (3) 1 (3)
b D z − (e0 a) βsin β −z + he + h p = −F31 − X33 φ,
he
2


 ∂z 2 


2 ∂x 2

where
 − h2e ( "  ! # )2
(1) 1
X33 = b ∈33 βsin β −z − he + hp dz,
− h2e −hp 2
(38)
 he
+hp ( "  ! # )2
(3) 2 1
X33 = b ∈33 βsin β −z + he + hp dz.
he
2
2
Summing Equations (32-a), (32-b) and (32-c) yields:
 (1) (3)

M1 + M2 + M3 = (e0 a)2 (M 001 + M 002 + M 003 ) + φ F31 + F31
!
1 00 02  (1) (2) (3)
 (39)
00
− W + W W D11 + D11 + D11 .
2
095122-9 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

By inserting Equation (31) in (39), the total bending moment is given:


!
1 00 02  
MT = F31 φ − D11 W + W W00
+ (e0 a)2 C Ẇ + mT Ẅ
2
x x  !0 (40)
2 0 2
+ (e0 a) W mT Ẇ 0 + W 0Ẅ 0 dxdx ,
l 0

where
MT = M1 + M2 + M3 ,
mtip
mT = mtot + H(x − l2 ), mtot = m1 + m2 + m3 ,
l2
(41)
(1) (3)
F31 = F31 + F31 ,
(1) (2) (3)
D11 = D11 + D11 + D11

Eventually, the first equation governing on the transverse vibrations of the nanobeam is obtained by
submitting Equation (40) in (30) and keeping terms up to the third order as:
!

00 3 0000 0 2
 1 02
0000
D11 W + W + 4W W W + W W 0 00 000
− 1 + W F31 φ 00 − W 0W 00F31 φ 0
2
 0
− W 0W 00(e0 a)2 C Ẇ + mT Ẅ + C Ẇ + mT Ẅ
x x  !0
2
+ W0 mT Ẇ 0 + W 0Ẅ 0 dxdx (42)
l 0
 x  x   ! 000
2 0 2 0 0
− (e0 a) W mT Ẇ 0 + W Ẅ dxdx
l 0
!
1 02   00
− 1 + W (e0 a)2 C Ẇ + mT Ẅ = 0.
2
After deriving the first governing equation, the electrical charge equation has to be derived. For that
purpose, Equations (33) and (34) are introduced into (21):
∂2 φ ∂2W
X11 − X33 φ − F31 2 = 0, (43)
∂x 2 ∂x
where
(1) (3)
X11 = X11 + X11 ,
(44)
(1) (3)
X33 = X33 + X33 ,
Because the clamped end of the nanobeam has a translational motion, i.e. W (0,t ) = W base (t), the
absolute transverse vibrations of the nanobeamcan be determined as below:
W (x, t) = Wrel (x, t) + Wbase (t), (45)
where W rel is the transverse displacement relative to the base and W base is the harmonic motion of
the base that is assumed in the following form:
Wbase = Ab sin(ωb t), (46)
where ωb and Ab are frequency and amplitude of the base vibrations, respectively. By applying
Equation (46), Equation (45) can be rewritten as:
W = Wrel + Ab sin(ωb t). (47)
Finally, equations of motion of a nanobeam subjected to a harmonic base excitation is obtained by
submitting Equation (47) in Equations (42) and (43):
095122-10 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)
!

0000 00 3 0 00 000 0000 0 2
 1 0 2
D11 Wrel + Wrel + 4Wrel Wrel Wrel + Wrel Wrel − 1 + Wrel F31 φ00
2
!
1 0 2   00
0
− Wrel 00
Wrel − 1 + Wrel (e0 a)2 C Ẇrel + mT Ẅrel
F31 φ0
2
x x  ! 000
2 0 0 2 0 0
− (e0 a) Wrel mT Ẇ rel + Wrel Ẅrel dxdx
l 0 (48)
 0
0
− Wrel 00
Wrel (e0 a)2 C Ẇrel + mT Ẅrel + C Ẇrel + mT Ẅrel
x x  !0
0 0 2 0 0
+ Wrel mT Ẇ rel + Wrel Ẅrel dxdx
l 0

= mT Ab ωb sin(ωb t) − CAb ωb cos(ωb t)


2

∂2 φ ∂ 2 Wrel
− X33 φ − F
X11
31 = 0. (49)
∂x 2 ∂x 2
By defining the following dimensionless parameters:
s  he
x Wrel φ A11 2
ξ= , w= , ϕ = , φ0 = , A11 = e
bC11 dz,
L he φ0 X113 h
− 2e

t he e0 a
τ= , η = , µ= , ω̄b = ωb t0 ,
t0 L L
F (1) φ0 F (3) φ0
r
m2 (1) (3) (1) (3)
t0 = L , F̄31 = 31 , F̄31 = 31 , F̄31 = F̄31 + F̄31 ,
A11 A11 he A11 he
(50)
(1) (2) (3)
(1) D11 (2) D11 (3) D11 (1) (2) (3)
D̄11 = , D̄11 = , D̄11 = , D̄11 = D̄11 + D̄11 + D̄11 ,
L 2 A11 L 2 A11 L 2 A11
(1) 2 (3) 2 (1) 2 2
(1) X11 φ0 (3) X11 φ0 (1) (3) (1) X33 φ0 L
X̄11 = , X̄11 = , X̄11 = X̄11 + X̄11 , X̄33 = ,
A11 he 2 A11 he 2 A11 he 2
(3) 2 2
(3) X33 φ0 L (1) (3) CL 2 CL
X̄33 = , X̄33 = X̄33 + X̄33 , C̄ = =√ .
A11 he 2 t0 A11 A11 m2
The nondimensional form of Equations (48) and (49) are given as:
η2
  !
D̄11 w 0000 + D̄11 η 2 w 003 + 4w 0 w 00 w 000 + w 0000 w 02 − F̄31 1 + w 02 ϕ 00 − F̄31 η 2 w 0 w 00 ϕ 0
2
00
η2
! !
mT
− µ2 1 + w 02 C̄ ẇ + ẅ
2 m2
"  ξ 
mT ξ  0 2
# 000 !0
 mT
−µ η w2 2 0 0 0
ẇ + w ẅ dξdξ − µ η w w C̄ ẇ +
2 2 0 00
ẅ (51)
1 m2 0 m2
"  ξ 
mT ξ  0 2
#0
mT 
+ C̄ ẇ + ẅ + η 2 w 0 ẇ + w 0 ẅ 0 dξdξ
m2 1 m2 0

mT Ab 2 Ab
= ω̄b sin (ω̄b τ) − C̄ ω̄b cos (ω̄b τ)
m2 h e he

∂2ϕ ∂2w
− X̄33 ϕ − F̄31 2 = 0.
X̄11 (52)
∂ξ2 ∂ξ
Where ω̄b is the dimensionless frequency of base excitation.
095122-11 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

V. DISCRETIZATION
In order to discretize the obtained partial differential equations (Equations (51) and (52)) and
reducing them to ordinary differential equations, assumed mode method is applied. In this method,
the transverse displacement w and the generated voltage ϕ are expressed as:
n
X
w (ξ, τ) = qi (τ) Pi (ξ), (53)
i=1
n
X
ϕ (ξ, τ) = si (τ) ri (ξ), (54)
i=1

where n is number of applied modes in these two expansions. Pi (ξ) and ri (ξ) are test functions
of transverse displacements and genereated voltage, respectively. The applied test functions have to
satisfy kinematic boundary conditions. qi (τ) and si (τ) are generalized coordinates that have to be
found. Pi (ξ) are considered as linear mode shapes of a cantilever beam:27
sinh (ψi ) − sin (ψi ) 
Pi (ξ) = cosh (ψi ξ) − cos (ψi ξ) − sinh (ψi ξ) − sin (ψi ξ) ,

(55)
cosh (ψi ) + cos (ψi )
where ψi are roots of the following characteristic equation:

cosh (ψi ) cos (ψi ) = −1. (56)

In order to satisfy the electrical charge boundary conditions, ri (ξ) is assumed as below:
" #
2i − 1
ri (ξ) = cos ( )πξ . (57)
2
In this research, only one mode has been employed. Therefore, subscripts in above equations will
be removed in the following equations. Now by introducing Equations (53) and (54) into Equation
(52) then multiplying the resulting equation by P (ξ) and integrating along the length of the beam, s
is obtained with respect to q:
1
∫ 0 F̄31 PP 00dξ
s=q 1 1
. (58)
∫ 0 X̄11 Pr 00dξ − ∫ 0 PX̄33 rdξ
Then by substituting Equations (53), (54) and (58) in (51) then multiplying by P (ξ) and integrating
along the length of the beam, an ordinary differential equation is obtained:

α1 q̈ + α2 q + α3 q3 + α4 qq̇2 + α5 q2 q̈ + α6 q̇ + α7 q2 q̇ = α8 ω̄b2 sin (ω̄b τ) + α9 ω̄b cos (ω̄b τ) , (59)

where
 1  1
mT 2 mT 00
α1 = P dξ − µ2 PP dξ
0 m2 0 m2
" 1 ! 1 ! #
mtip 00 l2 l2
− µ2 P2 H ξ − dξ + 2PP 0H0 ξ − dξ ,
m2 l2 0 l 0 l
 1 1 100 00
0000 ∫ F̄31 PP dξ ∫ 0 F̄31 Pr dξ
α2 = D̄11 PP dξ − 10 00 1
,
0 ∫ 0 X̄11 Pr dξ − ∫ 0 X̄33 P r dξ
 1 1 1
 1 ∫ F̄31 PP 00dξ ∫ 0 P 02 F̄31 Pr 00dξ
α3 = D̄11 η 2
P 003 P + 4PP 0P 00P 000 + PP 0000P 02 dξ − η 2 0 1 1
0 2 ∫ 0 X̄11 Pr 00dξ − ∫ 0 X̄33 P r dξ
1 1
∫ 0 F̄31 PP 00dξ ∫ 0 F̄31 PP 0P 00r 0dξ
− η2 1 1
, (60)
∫ 0 X̄11 Pr 00dξ − ∫ 0 X̄33 P r dξ
095122-12 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

 1 "  ξ   1 "  ξ 
mT ξ 0 2
# 0
mT ξ 0 2
# 000 
α4 = η 2 P P 0  dξ − µ2 η 2 P P 0
   
P dξdξ P dξdξ  dξ,
1 m2 0 1 m2 0

0 
 
0 

 
" 1 1 #
mT 02 00 mtip 0 2 00
α5 = α4 − µ2 η 2 PP P dξ + P P P dξ
0 m2 0 m2 l2

η2 2 1
(
mT 02 00
− µ PP P dξ
2 0 m2
" 1 ! 1 !#)
mtip 2 02 l2 02 l2
+ P P H 00
ξ− dξ + 2P PP H ξ −
0 0
,
m2 l 2 0 l 0 l
 1  1
α6 = C̄ P2 dξ − µ2 C̄ P P 00dξ,
0 0
 1
3
α7 = − µ2 η 2 C̄ P P 02 P 00dξ,
2 0
 1
mT Ab
α8 = P dξ,
0 m2 h e
 1
Ab
α9 = − PC̄ dξ.
0 he
Equation (59) can be rewritten as:
q̈ + β1 q + β2 q3 + β3 qq̇2 + β4 q2 q̈ + β5 q̇ + β6 q2 q̇ = β7 ω̄b 2 sin (ω̄b τ) + β8 ω̄b cos(ω̄b τ). (61)

VI. METHOD OF MULTIPLE SCALES


In order to plot the frequency-response and force-response diagrams of the nanoscale PEH,
multiple scale method has been employed. In order to apply this method, nonlinear terms have to be
balanced with the terms of damping and excitation by considering appropriate orders of ε for them.
For this purpose, Equation (61) is rewritten as:
q̈ + β1 q + β2 q3 + β3 qq̇2 + β4 q2 q̈ + ε 2 β5 q̇ + β6 q2 q̇ = ε 3 β7 ω̄b 2 sin (ω̄b τ) + ε 3 β8 ω̄b cos(ω̄b τ). (62)
q can be expanded as
q = εq1 (T0 , T2 ) + ε 3 q3 (T0 , T2 ) , (63)
where T 0 and T 2 are time scales that are defined as:
T0 = τ, T2 = ε 2 τ. (64)
It is worth mentioning that Equation (63) excluded q2 because there is no quadratic nonlinearity in
Equation (62). Regarding to the defined time scales, time derivatives are rearranged as:

d d2
= D0 + ε 2 D2 , = D0 2 + 2ε 2 D0 D2 . (65)
dτ dτ 2
The dimensionless frequency of the base excitation is considered to be close to the dimensionless
linear natural frequency of the nanobeam:
ω̄b = ω̄ + ε 2 σ, (66)
where ω̄ is dimensionless linear natural frequency and σ is detuning parameter. Substituting Equations
(63), (65) and (66) into (62) yields:
095122-13 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)
      3
D0 2 + 2ε 2 D0 D2 εq1 + ε 3 q3 + β1 εq1 + ε 3 q3 + β2 εq1 + ε 3 q3
  f  g2
+ β3 εq1 + ε 3 q3 D0 + ε 2 D2 εq1 + ε 3 q3
 2   
+ β4 εq1 + ε 3 q3 D0 2 + 2ε 2 D0 D2 εq1 + ε 3 q3
  
+ ε 2 β5 D0 + ε 2 D2 εq1 + ε 3 q3 (67)
 2   
+ β6 εq1 + ε 3 q3 D0 + ε 2 D2 εq1 + ε 3 q3
 2        
− ε 3 β7 ω̄ + ε 2 σ sin ω̄ + ε 2 σ T0 − ε 3 β8 ω̄ + ε 2 σ cos ω̄ + ε 2 σ T0

= 0.

By equating coefficients of ε 1 and ε 3 to zero, the following two equations are acquired:
ε 1 : D02 q1 + β1 q1 = 0, β1 = ω̄2 , (68)

ε 3 : D02 q3 + β1 q3 = −2D0 D2 q1 − β2 q13 − β3 q1 (D0 q1 )2 − β4 q12 D02 q1 − β5 D0 q1 − β6 q12 D0 q1


     (69)
+ β7 ω̄2 sin ω̄ + ε 2 σ T0 + β8 ω̄cos (ω̄ + ε 2 σ)T0 .

By solving linear Equation (68), q1 is given as:


q1 = A (T2 ) exp (iω̄T0 ) + Ā (T2 ) exp(−iω̄T0 ). (70)
By introducing Equation (70) into (69), then setting coefficients of secular terms, i.e. exp (iω̄T0 ), to
zero, solvability condition is obtained as:

−2iω̄A0 − 3 β2 A2 Ā − β3 A2 Āω̄2 + 3 β4 A2 Āω̄2 − i β5 Aω̄ − i β6 A2 Āω̄


1 1 (71)
+ β8 ω̄ exp (iσT2 ) − i β7 ω̄2 exp(iσT2 ) = 0.
2 2
Writing A in polar form as
1
A = P (T2 ) exp(iγ(T2 )) (72)
2
and then inserting it in Equation (71) yields:
! ! !
1 1 1 1
−2iω̄ P 0exp(iγ) + Piγ 0exp(iγ) − 3 β2 P2 exp(i2γ) P exp(−iγ)
2 2 4 2
! ! ! !
1 1 1 1
− β3 ω̄2 P2 exp(i2γ) P exp(−iγ) + 3 β4 ω̄2 P2 ei2γ P exp(−iγ)
4 2 4 2
! ! ! (73)
1 1 1
− i β5 ω̄ P exp(iγ) − iω̄ β6 P2 exp(i2γ) P exp(−iγ)
2 4 2
1 1
+ β8 ω̄ exp(iσT2 ) − i β7 ω̄2 exp(iσT2 ) = 0.
2 2
Collecting the coefficients of exp(iγ) in Equation (73) results in:
 3 1 3 1 1
−iω̄ P 0 + iPγ 0 − β2 P3 − β3 ω̄2 P3 + β4 ω̄2 P3 − Pω̄ β5 i − i β6 ω̄P3
8 8 8 2 8
(74)
1 1
− i β7 ω̄2 exp (iσT2 − iγ) + β8 ω̄ exp(iσT2 − iγ) = 0.
2 2
By defining
δ = σT2 − γ. (75)
095122-14 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

and employing Euler’s formula,


exp(iδ) = cos (δ) + i sin (δ) , (76)
Equation (74) can be rewritten as:
3 1 3 1 1
−iω̄ P 0 + iP σ − δ 0 β2 P3 − β3 ω̄2 P3 + β4 ω̄2 P3 − Pω̄ β5 i − i β6 ω̄P3


8 8 8 2 8 (77)
1 1
− i β7 ω̄2 (cos (δ) + i sin (δ)) + β8 ω̄ (cos (δ) + i sin (δ)) = 0.
2 2
By separating real and imaginary parts, modulating equations are obtained as:
1 1 1 1
P 0 = − P β5 − β6 P3 − β7 ω̄ cos (δ) + β8 sin (δ) , (78)
2 8 2 2
3 1 3 1 1
Pδ 0 = Pσ − β2 P3 − β3 ω̄P3 + β4 ω̄P3 − β7 ω̄ sin (δ) − β8 cos (δ) . (79)
8ω̄ 8 8 2 2
In order to achieve a steady state solution, the time derivatives have to be set to zero:
P 0 = Pδ 0 = 0, (80)
By inserting Equation (80) in Equations (78) and (79), the frequency-response equation is obtained:
!2 !2
1 1 1 1 3 1 3
− β7 2 ω̄2 − β8 2 + − P β5 − β6 P3 + Pσ − β2 P3 − β3 ω̄P3 + β4 ω̄P3 = 0, (81)
4 4 2 8 8ω̄ 8 8
where
σ = ω̄b − ω̄. (82)

VII. RESULTS AND DISCUSSION


In this section, the results of a numerical study on the proposed model are presented. Material
and geometric properties of the studied system is listed in Table I. First, the frequency-response curve
of the nanoscale PEH, obtained by a multiple scale method, is compared with a numeric method.
Then, frequency-response curves are investigated for different values of system parameters.
figure 2 shows comparison between frequency-response that is obtained by the multiple-scale
method and also a numeric method. The applied numeric method is the forth-order Runge-Kutta. In
that figure, a good agreement between the two applied solution methods can be seen. Furthermore,
stable and unstable branches are represented in figure 2.
The importance of considering the size effects in the model can also be manifested. Frequency-
response curves for different values of scale factor in absence of the tip mass is depicted in figure 3. An
interesting characteristic can be seen: For small values of the scale factor (i.e. lower than µ ≈ 0.165)
the nonlinear behavior of the nanobeam is of hardening type. For larger values of the scale factor

TABLE I. Material and geometric properties of the nanoscale PEH.26

Parameter Unit Numerical value


p
C11 GPa 132
e
C11 GPa 169
e31 Cm 2 -4.1
11 CV 1 m 1 5.841 × 10 9
33 CV 1 m 1 7.124 × 10 9
ρe Kg m 3 2330
ρp kg m 3 7500
L m 600 ×10 9
l1 m 0.6 × 10 9
b m 30 × 10 9
hp m 5 × 10 9
he m 5 × 10 9
095122-15 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

FIG. 2. Comparison between frequency-response curve obtained by a multiple-scale method and a numeric method (µ = 0,
mtip = 0, C = 3 × 10 6 and Ab = 2nm).

FIG. 3. Nonlinear frequency-response curves for different values of scale factor in absence of a tip mass (mtip = 0,
C = 3 × 10 6 and Ab = 2nm).

(larger than µ ≈ 0.165), nonlinear behavior is softening. Therefore, ignoring nonlocal effects results
in an incorrect predictions of a nanoscale PEH behavior. In addition, it is observed that increasing
the scale factor yields increases in both maximum vibration amplitude and generated voltage of the
PEH.
figure 4 displays the frequency-response curves for different values of the scale factor and tip
mass. It is observed that in presence of a tip mass, nanobeam demonstrates softening behavior and
this behavior is intensified by increasing the scale factor. Increasing the scale factor also increases
the maximum amplitude of the generated voltage. It should be noted that increasing the value of the
tip mass reduces the values of nonlinear natural frequencies more than linear natural frequencies of
the PEH. Therefore, the frequency-response curve bends to higher values of the detuning parameter.
The frequency-response curves for different values of the tip mass are demonstrated in figure 5.
It can be seen that increasing the value of the tip mass bends the frequency-response curve away
from σ = 0 to the left and increases the amplitude of the generated voltage in a broad band of
frequency.
After investigating the frequency–response curves, the effects of scale factor and excitation
frequency on the force response curves are studied. A Force-response curve determines the multiple-
valued regions for a known harmonic excitation. Multiple-valued region is a range of values of
095122-16 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

FIG. 4. Nonlinear frequency-response curves for different values of scale factor in presence of tip mass (mtip = 0.1 mtot ,
C = 3 × 10 6 and Ab = 2nm).

FIG. 5. Nonlinear frequency-response curve for different values of the tip mass ( µ = 0, C = 3 × 10 6 and Ab = 2nm).

FIG. 6. The force-response curve for different values of the scale factor (mtip = 0, C = 3 × 10 6 and σ = 3 × 10 4 ).
095122-17 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

FIG. 7. The force-response curve for different values of detuning parameter (mtip = 0, C = 3 × 10 6 and µ = 0).

excitation amplitude that multiple values of response amplitude are observed for each excitation
amplitude.
The force-response curves for different values of scale factor and detuning parameter are plotted
in figures 6 and 7, respectively. According to figure 6, by increasing the value of scale factor, the
jumps occur in higher values of the excitation amplitude. In figure 7, it can be observed that for some
values of the detuning parameter, there is no multiple-value region.

VIII. CONCLUSION
With the growing application of PEH in small electronic devices, preparing a more accurate
model for this type of energy harvesters is an essential task to maximize the harvested energy. For
this purpose, the nonlinear vibrations and generated voltage of a nanoscale PEH based on nonlocal
elasticity theory was studied in this research. Nanoscale PEH was modeled as a piezoelectric bimorph
cantilever nanobeam using Euler-Bernoulli beam theory. The nanobeam was assumed to be under a
harmonic base excitation. In order to consider the size effects in the prepared model, the nanlocal
elasticity theory as a strong nonclassic theory was applied. Nonlinear equations of motion were derived
and then solved using Hamilton’s principle and a multiple scales method, respectively. A numerical
method was also used to validate the obtained analytical solution. After solving the equations of
motion, Frequency-response and Force-response curves were plotted. Also, effects of scale factor
and tip mass on the generated voltage were investigated. Results showed the significant effect of the
scale factor in the nonlinear behavior of the nanoscale PEH. Therefore neglecting nonlocal effects
leads to an incorrect prediction of behavior of nanoscale PEH. Moreover, the results showed an
increase in the generated voltage and amplitude of vibrations by increasing the scale factor and also
nanobeam tip mass. To sum up, this paper provided a more accurate model for nanoscale piezoelectric
energy harvesters, and the presented modeling procedure could pave the way for modeling other types
of nanoscale vibration energy harvesters.

ACKNOWLEDGMENTS
The authors would like to thank Shahid Chamran University of Ahvaz for its financial support.
1 F. R. Fan, W. Tang, and Z. L. Wang, “Flexible nanogenerators for energy harvesting and self-powered electronics,” Advanced

Materials 28(22), 4283–4305 (2016).


2 T.Fan and L. Yang, “Surface effect on nano piezoelectric energy harvester based on flexural mode,” Polymer Composites
(2016).
3 A. Toprak and O. Tigli, “Piezoelectric energy harvesting: State-of-the-art and challenges,” Applied Physics Reviews 1(3),

031104 (2014).
4 X. Wang, “Piezoelectric nanogenerators—Harvesting ambient mechanical energy at the nanometer scale,” Nano Energy

1(1), 13–24 (2012).


095122-18 Foruzande, Hajnayeb, and Yaghootian AIP Advances 7, 095122 (2017)

5 X. Chen et al., “1.6 V nanogenerator for mechanical energy harvesting using PZT nanofibers,” Nano Letters 10(6),
2133–2137 (2010).
6 K. Wang and B. Wang, “An analytical model for nanoscale unimorph piezoelectric energy harvesters with flexoelectric

effect,” Composite Structures 153, 253–261 (2016).


7 K. Wang and B. Wang, “Non-linear flexoelectricity in energy harvesting,” International Journal of Engineering Science

116, 88–103 (2017).


8 Q. Deng et al., “Nanoscale flexoelectric energy harvesting,” International Journal of Solids and Structures 51(18), 3218–3225

(2014).
9 O. Aldraihem and A. Baz, “Energy harvester with a dynamic magnifier,” Journal of Intelligent Material Systems and

Structures 22(6), 521–530 (2011).


10 S. Ali, M. Friswell, and S. Adhikari, “Piezoelectric energy harvesting with parametric uncertainty,” Smart Materials and

Structures 19(10), 105010 (2010).


11 M. Fakhzan and A. G. Muthalif, “Harvesting vibration energy using piezoelectric material: Modeling, simulation and

experimental verifications,” Mechatronics 23(1), 61–66 (2013).


12 H. S. Kim, J.-H. Kim, and J. Kim, “A review of piezoelectric energy harvesting based on vibration,” International Journal

of Precision Engineering and Manufacturing 12(6), 1129–1141 (2011).


13 G. Wang, “Analysis of bimorph piezoelectric beam energy harvesters using Timoshenko and Euler–Bernoulli beam theory,”

Journal of Intelligent Material Systems and Structures 24(2), 226–239 (2013).


14 A. Erturk and D. J. Inman, “On mechanical modeling of cantilevered piezoelectric vibration energy harvesters,” Journal of

Intelligent Material Systems and Structures 19(11), 1311–1325 (2008).


15 F. Cottone et al., “Piezoelectric buckled beams for random vibration energy harvesting,” Smart Materials and Structures

21(3), 035021 (2012).


16 G. Sebald et al., “Experimental Duffing oscillator for broadband piezoelectric energy harvesting,” Smart Materials and

Structures 20(10), 102001 (2011).


17 L. Tang and Y. Yang, “A nonlinear piezoelectric energy harvester with magnetic oscillator,” Applied Physics Letters 101(9),

094102 (2012).
18 M. Eltaher, S. A. Emam, and F. Mahmoud, “Free vibration analysis of functionally graded size-dependent nanobeams,”

Applied Mathematics and Computation 218(14), 7406–7420 (2012).


19 S. Pradhan and T. Murmu, “Application of nonlocal elasticity and DQM in the flapwise bending vibration of a rotating

nanocantilever,” Physica E: Low-Dimensional Systems and Nanostructures 42(7), 1944–1949 (2010).


20 A. C. Eringen and D. Edelen, “On nonlocal elasticity,” International Journal of Engineering Science 10(3), 233–248 (1972).
21 L.-L. Ke and Y.-S. Wang, “Thermoelectric-mechanical vibration of piezoelectric nanobeams based on the nonlocal theory,”

Smart Materials and Structures 21(2), 025018 (2012).


22 L.-L. Ke, Y.-S. Wang, and Z.-D. Wang, “Nonlinear vibration of the piezoelectric nanobeams based on the nonlocal theory,”

Composite Structures 94(6), 2038–2047 (2012).


23 M. Nazemizadeh and F. Bakhtiari-Nejad, “Size-dependent free vibration of nano/microbeams with piezo-layered actuators,”

Micro & Nano Letters 10(2), 93–98 (2015).


24 Q. Wang, “On buckling of column structures with a pair of piezoelectric layers,” Engineering Structures 24(2), 199–205

(2002).
25 S. N. Mahmoodi and N. Jalili, “Non-linear vibrations and frequency response analysis of piezoelectrically driven

microcantilevers,” International Journal of Non-Linear Mechanics 42(4), 577–587 (2007).


26 A. H. Nayfeh and P. F. Pai, Linear and nonlinear structural mechanics. 2008: John Wiley & Sons.
27 D. J. Inman and R. C. Singh, Engineering vibration. Vol. 3. 2001: Prentice Hall Upper Saddle River.

You might also like