You are on page 1of 8

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO.

5, OCTOBER 2013

1527

A Piezoelectric Energy Harvester for Rotary Motion


Applications: Design and Experiments
Farbod Khameneifar, Siamak Arzanpour, and Mehrdad Moallem

AbstractThis paper investigates the analysis and design of a


vibration-based energy harvester for rotary motion applications.
The energy harvester consists of a cantilever beam with a tip
mass and a piezoelectric ceramic attached along the beam that
is mounted on a rotating shaft. Using this system, mechanical vibration energy is induced in the flexible beam due to the gravitational force applied to the tip mass while the hub is rotating. The
piezoelectric transducer is used to convert the induced mechanical vibration energy into electricity. The equations of motion of
the flexible structure are utilized along with the physical characteristics of the piezoelectric transducer to derive expressions for
the electrical power. Furthermore, expressions for the optimum
load resistance and maximum output power are obtained and validated experimentally using PVDF and PZT transducers. The results indicate that a maximum power of 6.4 mW at a shaft speed
of 138 rad/s can be extracted by using a PZT transducer with dimensions 50.8 mm 38.1 mm 0.13 mm. This amount of power
is sufficient to provide power for typical wireless sensors such as
accelerometers and strain gauges.
Index TermsCantilever beam, energy harvesting, piezoelectric
transducers, power optimization, rotational motion.

I. INTRODUCTION
EAL-TIME condition monitoring of rotating machines
and structures such as turbines and tires is highly desirable to achieve improved safety and health monitoring. With
advancements in wireless technology, modern fault detection
mechanisms for rotary applications can be implemented by installing wireless sensors on the structure to transmit data to a
health and status monitoring unit. Wireless systems do not have
the drawback of rotating wired systems that require the use of
slip rings to transfer sensory data. However, wireless sensors
and their transmission units need batteries as power sources for
their operation [1].

Manuscript received June 13, 2011; revised September 26, 2011,


January 26, 2012, and May 5, 2012; accepted May 19, 2012. Date of publication July 12, 2012; date of current version July 11, 2013. Recommended by
Technical Editor J. Wang. This work was supported in part by the Natural
Sciences and Engineering Research Council of Canada (NSERC) under the
Discovery Grants program, in part by the Simon Fraser University Start-up
fund, and in part by the British Columbia NRAS fund.
F. Khameneifar was with the Department of Mechatronic Systems Engineering, School of Engineering Science, Simon Fraser University, Surrey, BC
V3T 0A3, Canada. He is now with the Department of Mechanical Engineering, University of British Columbia, Vancouver, BC V6T 1Z4, Canada (e-mail:
farbodkh@interchange.ubc.ca).
S. Arzanpour and M. Moallem are with the Department of Mechatronic
Systems Engineering, School of Engineering Science, Simon Fraser University,
Surrey, BC V3T 0A3, Canada (e-mails: arzanpour@sfu.ca; mmoallem@sfu.ca).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TMECH.2012.2205266

A major shortcoming of using batteries is the requirement to


recharge or replace them on a regular basis [2]. Since stopping
the systems to replace the batteries may not be practical, an
apparatus such as an energy harvester that can locally scavenge
or harvest energy to recharge the batteries is highly advantageous. Several energy harvesting mechanisms have been developed for transforming ambient energy (in this case, mechanical
vibrations) into electricity. Among them, electrostatic, electromagnetic, and piezoelectric harvesters have been proposed [3].
Electrostatic generators require an additional power source to
operate, whereas electromagnetic transducers generate low output voltages. Roundy et al. [4] compared the aforementioned
energy harvesting techniques and concluded that piezoelectric
transducers are good candidates for converting vibration energy
into electricity. As noted in [5], the energy density of piezoelectric transducers is three times higher when compared to electrostatic and electromagnetic transducers. A piezoelectric energy
harvester is usually comprised of a cantilever beam on which a
piezoceramic layer and a tip mass are mounted. Sodano et al.
[6] tested three types of piezoelectric devices, i.e., the monolithic piezoceramic material LeadZirconateTitanate (PZT),
the Quick Pack (QP), and the macrofiber composite (MFC), and
compared their capability to charge a battery. Their investigation indicated that MFC is not well suited for energy harvesting,
whereas QP and PZT are both capable of efficiently recharging
batteries. Different models have been utilized to describe the
electromechanical behavior of piezoelectric energy harvesters.
Sodano et al. [7] used the RayleighRitz solution for modeling
a piezoelectric energy harvester beam without a tip mass. Erturk
and Inman [8] concluded that a lumped model may yield inaccurate results for predicting the motion of cantilever beams and
proposed a coupled distributed solution [9], [10]. The placement
of PZT electrodes to achieve efficient energy harvesting was investigated by Erturk et al. [11] through a study of the strain nodes
of a cantilever harvester. Wickenheiser et al. [12] studied the effects of electromechanical coupling of a cantilever harvester on
the dynamics of charging a storage capacitor. Other interesting applications of piezoelectric energy harvesting include the
works by Almouahed et al. [13], and Platt et al. [14], [15], who
used piezoelectric energy harvesters to provide power for in vivo
knee implants.
Most energy harvesters that have been reported so far in the
literature rely on the vibrations induced from a base excitation. In the energy harvester discussed in this paper, a novel
method has been used to excite the beam vibration. We utilize
the gravitational force on the tip mass to generate continuous
oscillations in a cantilever beam during its rotating motion. In
Section II, a simplified single-degree-of-freedom model of the

1083-4435/$31.00 2012 IEEE

1528

Fig. 1.

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 5, OCTOBER 2013

Schematic view of the energy harvester mounted on a rotating hub.

piezoelectric energy harvesting system is presented. Hence, the


dynamic equations describing the interaction of the piezoelectric
transducer and flexible beam are obtained. Expressions describing the electrical output voltage and power of the combined
piezoelectric energy harvesting system are presented including
the optimal value of load resistance connected to the energy
harvester. Furthermore, an upper bound for the output power
is obtained through impedance matching for the proposed design. In Section III, details of the test bed used and experimental results are presented. The results indicate close agreement
between theory and experiments including system dynamics,
generated output voltage of the harvester, and the optimal load
for maximizing the output power.
II. ROTATING HARVESTER DESIGN AND MATHEMATICAL
MODEL
The proposed energy harvester consists of a cantilever beam
with a piezoelectric layer carrying a tip mass with the whole
system mounted on a rotating hub. A schematic diagram of the
mechanism is shown in Fig. 1. The dimensions of the beam
structure in Fig. 1 are provided in Table III. As shown in the
figure, the hub is rotating about the Z axis and gravity is acting
along the negative Y axis.
The main source of vibration input to the harvester is the
alternating gravitational force on the cantilever beam. The concept is shown in Fig. 2 for one cycle of rotation of one of the
four cantilever beams. When the beam is in position (a), the
gravitation force on the tip mass applies a bending moment
on the cantilever beam in the negative Z direction. As the beam
reaches position (b), the bending moment on the cantilever beam
is zero. At position (c), the bending moment due to the tip mass
is in the positive Z direction. Position (d) is identical to (b),
where the effect of the tip-mass gravitational force is zero. As a

Fig. 2. Orientation of the beam and gravitational force in one cycle of rotation
and the material mode orientation.

result, a 360 rotation of the hub results in the application of an


alternating force on the harvester. The frequency of this force
is as same as the hub rotational frequency. Thus, the induced
vibration generates a harmonic voltage on the piezoelectric element that is a function of the strain applied to this element. The
material mode orientation of the PZT transducer is also shown
in Fig. 2(a), where X and Y axes correspond to dimensions
1 and 3, respectively.
Detailed modeling of the dynamics of the rotating energy
harvester with tip mass is presented in [16]. In the following, we
provide a brief overview of the derivation of dynamic equations
using a single mode analysis.
A. Vibration Model of the Rotating Beam With a Piezoelectric
Element
Let us consider the EulerBernoulli beam equation that describes the vibration behavior of a cantilever beam as follows
(see e.g., [10]):
EI

2 w(, t)
4 w(, t)
+ AL4
=0
4

t2

(1)

where the deflection of the beam relative to its base is denoted


by w(, t), = x/L is the normalized position x along the beam
with length L, E is Youngs modulus, I is the area moment of
inertia, is the mass density, and A is the cross-sectional area
of the beam. Here, E, I, , and A are not functions of due to
the uniform distribution of the piezoelectric layers. Moreover,
placing the piezoelectric layers on a part of the beam would not
be appropriate if the device is to be used as an energy harvester
as the goal is to extract maximum power. The solution of (1)

KHAMENEIFAR et al.: PIEZOELECTRIC ENERGY HARVESTER FOR ROTARY MOTION APPLICATIONS: DESIGN AND EXPERIMENTS

can be obtained by using the separation of variables as follows:


w(, t) = ().(t)

(2)

where () is the shape eigenfunction and (t) is the modal


mechanical response. Next, we utilize the Lagrangian formulation to derive a mathematical model that describes the system
dynamics, including the piezoelectric beam, tip mass, and the
rotating hub. The kinetic and potential energies of the system,
T and U, are respectively given as follows:

1
1 2 1

(w 2 + 2wx
)dm
+ Jh 2
T = Jb +
2
2 b eam
2

2
1
1
w(L,

t)
+ JL +
+ ML (2 w(L, t)2
2
x
2
2)
+ (w(L,

t) + L)
2

 L
EI 2 w
dx + gA
U=
2
x2
0
 L

w

+ (x w tan ) sin dx
cos
0


w| =1
+ (L w| =1 tan ) sin
+ ML g
cos

(3)

(4)

where M22 is the symmetric inertia matrix and C22 is the


vector of Coriolis and centrifugal forces. The induced moment
FP (t) is given by [17]
(6)

The elements of the matrices M and C are given by the following terms:
m11 = Jh + Jb + JL + ML L2 + ML (e )2
m12 = m21 = ML Le +
m22 =

mb 2e

ML 2e

JL e


JL e2

g11



= gA( sin )

d + ML g( sin )e

L2
cos + ML gL cos
2
  1

= g cos A
d + ML e
+ gA

g21

(13)
(14)

where mb is the beam mass, subscript e denotes the end of the


beam, and is a function of the integral of mode shape given
1
by = AL2 0 ()d.
Furthermore, the equivalent spring constant K in (5) is given
by

  1
2
d
(15)
K = EI
0

where ML and JL are the load mass and inertia, and Jb and Jh
are beam and hub inertia, respectively. Let us define the vector
= [ ]T with the vector of
of generalized coordinates as q
= [ FP ]T , where is the
generalized force defined as F
angular displacement of the hub [16], is the applied torque
to the hub, and FP is the moment induced by the piezoceramic
layer. The dynamic equations of the flexible beam can then be
obtained by using the Lagrangian formulation as follows:
 
 

M (, t)
+ C(, , , , t)
+ G(, , t)



 
0
(t)
(5)
+
=
K(t)
FP (t)

FP (t) = kv(t)[ (0)  (1)].

follows:

1529

(7)

where is the backward coupling term. The coupling term is a


function of geometric parameters of the system, the piezoelectric
constant, and the modulus of elasticity of the piezoelectric layer
and the flexible beam. Since the objective of this study is to
harvest energy at the steady-state speed of a rotary system,
the angular acceleration is set to zero in (5), i.e., = 0. The
effect of damping can be added to the EulerBernoulli equation
by incorporating internal strain rate damping and viscous air
damping.
Therefore, the modal damping coefficient 1 can be expressed
C 2
by 21 1 = sE 1 + Cma , where the first term represents the effect of the strain-rate damping and the second term represents
air damping. Other parameters represent the damping coefficients Cs , and Ca , and the natural frequency of the first mode
1 respectively [16]. Incorporating all the terms in (5) results in
+ 21 1 (t)
+ 2 (t) + X v(t) = A cos t
(t)
1
B
B

(16)

where X, A, and B are given by


X = e
 
A = A

(17)

d + ML e g

(18)

0
2
B = (ML 2e + JL 2
e + mb e ).

The natural frequency of the first mode 1 in (16) is


where C is given by
C = (K ML 2 2e ).

(19)

C/B
(20)

(8)
(9)

c11 = ML 2e

(10)

c12 = c21 = ML 2e

(11)

c22 = 0

(12)

where subscript e denotes that the parameter is evaluated at


the tip of the beam. The elements of the vector G(, , t) are as

B. Piezoelectric Transducer Electrical Model


A piezoelectric element under excitation can be modeled as
a current source in parallel with an internal capacitance [18].
Fig. 3 shows the circuit representation of a piezoelectric element
connected across a purely resistive load, which represents the
effect of an energy storage, or energy consuming load.
The electrical circuit equation of the system, considering the
first mode of vibrations, can be obtained by using Kirchhoffs

1530

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 5, OCTOBER 2013

Fig. 3. Electrical circuit symbolizing the resistive load connected to the single
piezoelectric layer.

laws as follows [17]:


+
CP v(t)

v(t)
d31

= E hpn be (t)


Rl
S11

(21)

where Rl is the load resistance; Cp = s33 bL/hp is the internal capacitance of piezoelectric layer; = Sd 3E1 hpn be is the
11
forward coupling term; s33 is the permittivity constant; b, L,
and hp are the width, length, and thickness of the piezoceramic
layer, respectively; hpn is the distance between the neutral axis
and the center of piezoceramic layer; d31 is the piezoelectric
E
is the elastic compliance at a constant electric
constant; and S11
field. The right-hand side of (21) represents ip (t). Referring to
Fig. 2(a), there are two operating modes, d31 and d33 , through
which the piezoelectric transducer can generate electricity. In
the d33 -mode, both the mechanical stress and output voltage act
along dimension 3. In the d31 -mode, the mechanical stress acts
along dimension 1, while the voltage acts along dimension 3. As
discussed in previous works (e.g., [19], [20]), a thin piezoelectric layer bonded to a substrate cantilever beam that operates
in the d31 -mode can produce larger strains with smaller input
forces.

The amplitude of the vibration can be obtained from (16)


and (21) to yield the amplitude of output voltage given by


l A
2R

(22)
V () =
2
D1 + D22
where the D1 and D2 terms are given by



C
+ ML 2e 2 K (23)
D1 = B + 2Rl CP B 2
B

2Rl (X + CP K) + 2B

C
B

(24)

Using (22), the average harvested electrical power is


presented by
|P | =

|V |2
42 Rl 2 A2
.
=
Rl
(D12 + D22 )

Experimental setup used for validating the analytical model.

TABLE I
GEOMETRIC AND PHYSICAL PARAMETERS OF THE PVDF ENERGY HARVESTER

III. EXPERIMENTAL VALIDATION

C. Closed-Form Solution of the Electrical Output Power

D2 = 2Rl CP (B + ML 2e )3

Fig. 4.

(25)

Differentiating (25) with respect to Rl (resistive load), an optimal load to maximize the amplitude of harvested power can be
obtained. An important issue that should be noted here is the effect of damping on the optimal resistive load. By differentiating
the expression for the optimal resistance in terms of the damping
ratio, it can be concluded that an increase in the damping term
will result in an increase in the value of the optimal resistance.

A. Experimental Setup for Vibration Energy Harvesting From


a Rotating Hub
The experimental setup for generating electric voltage from
a rotating hub is shown in Fig. 4. Two different energy harvesters were utilized in our tests, i.e., a PZT manufactured by
MIDE Inc., and a PVDF film manufactured by Images SI, Inc.
(PZ-03). The harvesters are attached to a flexible cantilever
beam using Loctite glue (Model No. 330). A tip mass is also attached to the cantilever beam (48 and 65 g) for tuning the natural
frequency of the structure. The geometric, physical, and material properties of the piezoelectric layer and substructure are
given for the PVDF and PZT harvesters in Tables I and II, and
Tables III and IV, respectively. The cross hub has room for four
harvesters; however, for proof of concept demonstration only
one is utilized. The output power can increase by using the full
capacity of the system and employing a suitable power management unit. The shaft is driven by a dc motor from Maxon
Motors (A-max 32 Model No. 236669) equipped with a shaft
encoder with 500 Counts per (HEDL 5540 Model No. 110514).
The maximum power in this harvester occurs at its natural frequency with an optimal load. To identify the natural frequency
the hub velocity is gradually increase from 0 to 150 rad/s and the
output voltage is measured for an arbitrary resistive load (100 ).
The angular velocity at which the maximum voltage is achieved
represents the natural frequency of the beam. The optimal resistive load was then identified by maintaining the hub angular

KHAMENEIFAR et al.: PIEZOELECTRIC ENERGY HARVESTER FOR ROTARY MOTION APPLICATIONS: DESIGN AND EXPERIMENTS

1531

TABLE II
MATERIAL PARAMETERS OF THE PVDF ENERGY HARVESTER

Fig. 5. Output power versus resistive load for 48-g tip mass: PVDF experimental data (blue dotted line) and theoretical data (red line).

TABLE III
GEOMETRIC AND PHYSICAL PARAMETERS OF THE PZT ENERGY HARVESTER

Fig. 6. Output power versus resistive load for 65-g tip mass for PVDF: experimental data (blue dotted line) and theoretical data (red line).

TABLE IV
MATERIAL PARAMETERS OF THE PZT ENERGY HARVESTER

Fig. 7. Output voltage for 48-g tip mass and R = 600 k: experimental data
(blue dotted line) and theoretical data (red line).

B. Validation of the Single-Mode Closed Form Expressions for


the Output Electric Voltage
velocity at the harvester natural frequency and varying the load
resistance until a maximum in the power was achieved. The
results for the power versus resistive load are shown in Figs. 5
and 6 for different tip masses. Figs. 7 and 8 illustrate the output
voltage of the harvester at the optimal load for different hub
velocities.

The mathematical model in (22) and (25) are used to calculate the output voltage, optimal resistive load, and maximum
power. The strain rate damping in those equations is obtained
experimentally using the exponential decay of the response of
the beam in an impact hammer test. The damping term due to air
flow was not considered in our simulations. The output voltage

1532

Fig. 8. Output voltage for 65-g tip mass and R = 600 k: experimental data
(blue dotted line) and theoretical data (red line).

of the harvester with a 600-k resistive load was simulated for


the 48 and 65 g tip masses at different hub angular velocities as
depicted in Figs. 7 and 8, respectively. The results indicate close
agreement between the theory and experiments. There exists
a small discrepancy between the experimental and simulation
results, i.e., 8.6% for voltage response of the PVDF harvester
with a 48-g tip mass, and 5.8% for a 65-g tip mass. Also, the
simulated optimal resistive load of the harvester for the 48 and
65-g tip masses were 560 and 540 k, respectively, i.e., corresponding to 6.7% and 10% errors, respectively. Figs. 5 and
6 illustrate the experimental and simulation results of the harvested power versus different resistive loads. The largest error
occurs close to the natural frequency where the beam vibration
and consequently the air damping effect is maximum.
To verify the effect of imbalance of the shaft on the output
voltage, we performed measurements when the setup was oriented in the vertical position where the gravity cannot act as the
source of beam excitation. The output voltage that was mainly
due to the shaft imbalance induced vibrations was less than 2%
of the horizontal configuration. This experiment indicates that
the effect of imbalance can be neglected.
The experimental and simulation results indicate that the effect of tip mass on the optimum load resistance is negligible.
Although the changes in the tip mass intuitively alter the damping ratio the value of viscous damping coefficient remains
the same. The experimental output power extracted from this
device using the 65-g tip mass at the optimal resistive load was
30.8 W at 85 rad/s (natural frequency).
Similar experiments were conducted for measuring the output
voltage and calculating the optimal resistive load using the PZT
layer. It should be noted that the PZT has a higher modulus of
elasticity and thus a higher natural frequency. In the setup used in
this study, the angular velocity of the hub cannot reach velocities
beyond 200 rad/s due to the limitation on the maximum speed
of the dc motor. To reduce the natural frequency of the beam a
105-g tip mass was used. The natural frequency of the harvester beam with this tip-mass is 138 rad/s. A simulation was
conducted on the PZT harvester with the specifications summarized in Tables III and IV. The same approach used for the
PVDF was used here.

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 5, OCTOBER 2013

Fig. 9. Output voltage of PZT for 105-g tip mass and R = 40 k: experimental
data (blue dotted line) and theoretical data (red line).

Fig. 10. Output power versus load resistance for PZT for a 105-g tip mass:
experimental data (blue dotted line) and theoretical data (red line).

Fig. 9 shows the experimental and theoretical results for the


voltage output of the PZT energy harvester for an optimal load
of 40 k, which was obtained from the power versus load resistance plot. The maximum error of voltage at the resonance
frequency is 10.3% for the PZT harvester.
The output power versus resistive load is plotted for the PZT
harvester as shown in Fig. 10. Based on the experimental results,
the maximum extracted power of PZT energy harvester occurred
when the load resistance is 40 k. This number matches with
the resistive load calculated from the mathematical model. The
experimentally measured maximum output power from the PZT
energy harvester is 6.4 mW, which is very close to the value
calculated from (25).
To be able to compare the extracted power from the PVDF
harvester and the PZT harvester, one should consider similar
physical conditions in terms of the tip mass and length of the
beam. For the PVDF harvester with a 105-g tip mass, a maximum output power of 147 W was obtained when the harvester
was connected to the optimal resistance. Comparing the results of the two harvesters investigated in this study, it may be
concluded that the PVDF transducer has a larger optimal resistive load than the PZT transducer (600 k in comparison with
40 k). The results indicate that the amplitude of generated
power from the PZT harvester is about 44 times higher than the
case for PVDF (6.4 mW compared with 147 W, when using
the same lengths and tip masses). This result is in agreement

KHAMENEIFAR et al.: PIEZOELECTRIC ENERGY HARVESTER FOR ROTARY MOTION APPLICATIONS: DESIGN AND EXPERIMENTS

with previous studies indicating that the optimal load resistance


is higher for PVDF case due to the low piezoelectric constant
of these materials [21]. The amount of power obtained from
the PZT harvester beam (6.4 mW) with a tip mass of 105 g is
enough for supplying a typical wireless sensor [22]. The power
harvested using the PVDF with the same dimensions and a tip
mass of 105 g was 147 W, which can only be used to power
a wireless sensor in the sleeping state which requires a typical
power in the range 103 to 101 mW [22]. The output power
from the PVDF harvester beam with lower tip masses of 48 and
65 g, were 10.4 and 30.8 W, respectively. These power levels
are not suitable for powering a wireless sensor in the sleeping
state. The PZT energy harvester may thus be a better candidate
when compared to the PVDF harvester due to its higher output
power. However, in applications where variation of the mechanical torque to the rotating hub is large, the risk of failure in the
PZT harvester is higher than that of the PVDF since PZT is brittle and PVDF is highly flexible. Further studies involving the
utilization of impedance concept for power optimization have
been reported by Liang et al. [23].
IV. CONCLUSION
A novel piezoelectric energy harvester for rotary motion applications was presented in this study. The piezoelectric energy harvester consists of a cantilever beam and a tip mass
mounted on a rotating hub. When the hub rotates with a constant angular velocity, the gravity force on the tip mass causes
the mass-beam system to vibrate. The steady-state closed-form
electromechanical expressions were used to find the maximum
harvested power for the optimal load resistance. Experimental
studies confirm analytical predictions in terms of vibration response, output voltage, optimal load of the harvester, and output
power extracted from the energy harvester. Two different energy
harvesters were tested; one using a PVDF film and another using a PZT transducer. The results indicate that the output power,
when a PZT transducer is used, is about 44 times higher than the
case when a PVDF film is used. As a result, the PZT harvester
is a good candidate to be utilized as a local miniaturized power
generator for wireless sensors in applications involving rotary
motion condition monitoring. Using more than one harvester
on the hub would contribute to the generation of more power;
however, further experimental studies need to be conducted to
evaluate power generation capability of the system.

1533

[6] H. A. Sodano, D. J. Inman, and G. Park, Comparison of piezoelectric


energy harvesting devices for recharging batteries, J. Intell. Mater. Syst.
Struct., vol. 16, pp. 799807, 2005.
[7] H. A. Sodano, G. Park, and D. J. Inman, Estimation of electric charge
output for piezoelectric energy harvesting, Strain J., vol. 40, pp. 4958,
2004.
[8] A. Erturk and D. J. Inman, On mechanical modeling of cantilevered
piezoelectric vibration energy harvesters, J. Intell. Mater. Syst. Struct.,
vol. 19, pp. 13111325, 2008.
[9] A. Erturk and D. J. Inman, Issues in mathematical modeling of piezoelectric energy harvesters, Smart Mater. Struct., vol. 17, pp. 114, 2008.
[10] A. Erturk and D. J. Inman, An experimentally validated bimorph cantilever model for piezoelectric energy harvesting from base excitations,
Smart Mater. Struct., vol. 18, pp. 118, 2009.
[11] A. Erturk, P. A. Tarazaga, J. R. Farmer, and D. J. Inman, Effect of strain
nodes and electrode configuration on piezoelectric energy harvesting from
cantilevered beams, ASME J. Vib. Acoust., vol. 131, pp. 111, 2009.
[12] A. M. Wickenheiser, T. Reissman, W. Wu, and E. Garcia, Modeling the
effects of electromechanical coupling on energy storage through piezoelectric energy harvesting, IEEE/ASME Trans. Mechatronics, vol. 15,
no. 3, pp. 400411, Jun. 2010.
[13] S. Almouahed, M. Gouriou, C. Hamitouche, E. Stindel, and C. Roux, The
use of piezoceramics as electrical energy harvesters within instrumented
knee implant during walking, IEEE/ASME Trans. Mechatronics, vol. 16,
no. 5, pp. 799807, Oct. 2011.
[14] S. R. Platt, S. Farritor, and H. Haider, On low-frequency electric
power generation with PZT ceramics, IEEE/ASME Trans. Mechatronics,
vol. 10, no. 2, pp. 240252, Apr. 2005.
[15] S. R. Platt, S. Farritor, K. Garvin, and H. Haider, The use of piezoelectric ceramics for electric power generation within orthopedic implants,
IEEE/ASME Trans. Mechatronics, vol. 10, no. 4, pp. 455461, Aug. 2005.
[16] F. Khameneifar, M. Moallem, and S. Arzanpour, Modeling and analysis
of a piezoelectric energy scavenger for rotary motion applications, ASME
J. Vib. Acoust., vol. 133, pp. 16, 2011.
[17] M. R. Kermani, M. Moallem, and R. V. Patel, Parameter selection and
control design for vibration suppression using piezoelectric transducers,
Control Eng. Pract., vol. 12, pp. 10051015, 2004.
[18] G. K. Ottman, H. F. Hofmann, A. C. Bhatt, and G. A. Lesieutre, Adaptive
piezoelectric energy harvesting circuit for wireless remote power supply,
IEEE Trans. Power Electron., vol. 17, no. 5, pp. 669676, Sep. 2002.
[19] N. E. duToit, B. L. Wardle, and S. G. Kim, Design considerations
for MEMS-scale piezoelectric mechanical vibration energy harvesters,
Integr. Ferroelectr., vol. 71, pp. 121160, 2005.
[20] A. Hande, R. Bridgelall, and D. Bhatia, Energy harvesting for active
RF sensors and ID tags, in Energy Harvesting Technologies. Berlin,
Germany: Springer, 2008, ch. 18.
[21] D. Shen, S. Y. Choe, and D. J. Kim, Comparison of piezoelectric materials
for vibration energy conversion devices, in Proc. Mater. Res. Soc. Symp.,
2006, vol. 966, pp. 16.
[22] D. Steingart, Power sources for wireless sensor networks, in Energy
Harvesting Technologies. Berlin, Germany: Springer, 2008, ch. 9.
[23] J. Liang and W. Liao, Impedance modeling and analysis for piezoelectric
energy harvesting systems, IEEE/ASME Trans. Mechatronics, to be
published.

REFERENCES
[1] W. Wang, F. Ismail, and F. Golnaraghi, A neuro-fuzzy approach to gear
system monitoring, IEEE Trans. Fuzzy Syst., vol. 12, no. 5, pp. 710723,
Oct. 2004.
[2] M. Bhardwaj, T. Garnett, and A. P. Chandrakasan, Upper bounds on the
lifetime of sensor networks, in Proc. IEEE Int. Conf. Commun., vol. 3,
2001, pp. 785790.
[3] C. B. Williams and R. B. Yates, Analysis of a micro-electric generator
for microsystems, Sens. Actuators A, vol. 52, pp. 811, 1996.
[4] S. Roundy, P. Wright, and J. Rabaey, A study of low level vibrations as
a power source for wireless sensor nodes, Comput. Commun., vol. 26,
pp. 11311144, 2003.
[5] S. Priya, Advances in energy harvesting using low profile piezoelectric
transducers, J. Electro-Ceram., vol. 19, pp. 165182, 2007.

Farbod Khameneifar received the B.Sc. degree in


mechanical engineering from the University of
Tehran, Tehran, Iran, in 2008, and the M.Sc. degree
in mechatronic systems engineering from Simon
Fraser University, Surrey, BC, Canada, in 2011.
He is currently working toward the Ph.D. degree in
mechanical engineering at the University of British
Columbia, Vancouver, BC, Canada.
His current research interests include computeraided design, manufacturing and inspection
(CAD/CAM/CAI), computational geometry, geometric and dynamic modeling, and design and optimization of mechatronic
systems.

1534

Siamak Arzanpour received the B.Sc. degree from


the University of Tehran, Tehran, Iran, in 1998,
the M.Sc. degree from the University of Toronto,
Toronto, ON, Canada, in 2003, and the Ph.D. degree from the University of Waterloo, Waterloo, ON,
Canada, all in mechanical engineering.
He is currently an Assistant Professor at Simon
Fraser University, Surrey, BC, Canada. His current
research interests include a wide range of topics, including smart materials, vibration, haptic systems,
pattern and material recognition using vibration signatures of biomaterials, and energy harvesting from mechanical vibrations for
remote sensors.

IEEE/ASME TRANSACTIONS ON MECHATRONICS, VOL. 18, NO. 5, OCTOBER 2013

Mehrdad Moallem received the B.Sc. degree from


Shiraz University, Shiraz, Iran, in 1986, the M.Sc. degree from Sharif University of Technology, Tehran,
Iran, in 1988, both in electrical and electronic engineering, and the Ph.D. degree in electrical and computer engineering from Concordia University, Montreal, QC, Canada, in 1997.
From 1998 to 1999, he was a Research and Development Engineer at Duke University, Durham, NC.
From 1999 to 2007, he was an Assistant, and then an
Associate Professor in the Department of Electrical
and Computer Engineering, The University of Western Ontario, London, ON,
Canada. Since 2007, he has been with the Mechatronics Systems Engineering
Program, School of Engineering Science, Simon Fraser University, Surrey, BC,
Canada, where he is currently a Professor in the Faculty of Applied Sciences.
His current research interests include control applications, in particular, vibration control, power electronic control for energy conversion, smart sensors and
actuators, and embedded real-time computing. He has authored or coauthored
extensively in the aforementioned areas and is the coauthor of four technical
books on vibration control using piezoelectric transducers, control of flexible
robots, medical robotics, and wind energy conversion.
Dr. Moallem has served on the Editorial Boards of several conferences
and journals including the American Control Conference and the IEEE/ASME
TRANSACTIONS ON MECHATRONICS.

You might also like