• 沒有找到結果。

An improved analysis of the SSHI interface in piezoelectric energy harvesting

N/A
N/A
Protected

Academic year: 2021

Share "An improved analysis of the SSHI interface in piezoelectric energy harvesting"

Copied!
13
0
0

加載中.... (立即查看全文)

全文

(1)

An improved analysis of the SSHI interface in piezoelectric energy harvesting

This article has been downloaded from IOPscience. Please scroll down to see the full text article.

2007 Smart Mater. Struct. 16 2253

(http://iopscience.iop.org/0964-1726/16/6/028)

Download details:

IP Address: 140.112.113.225

The article was downloaded on 22/12/2008 at 01:30

Please note that terms and conditions apply.

The Table of Contents and more related content is available

(2)

Smart Mater. Struct. 16 (2007) 2253–2264 doi:10.1088/0964-1726/16/6/028

An improved analysis of the SSHI

interface in piezoelectric energy harvesting

Y C Shu

1,3

, I C Lien

1

and W J Wu

2

1Institute of Applied Mechanics, National Taiwan University, Taipei 106, Taiwan, Republic of China

2Department of Engineering Science and Ocean Engineering, National Taiwan University, Taipei 106, Taiwan, Republic of China

E-mail:yichung@spring.iam.ntu.edu.tw

Received 29 April 2007, in final form 15 August 2007

Published 9 October 2007

Online at

stacks.iop.org/SMS/16/2253

Abstract

This paper provides an analysis for the performance evaluation of a

piezoelectric energy harvesting system using the synchronized switch

harvesting on inductor (SSHI) electronic interface. In contrast with estimates

based on a variety of approximations in the literature, an analytic expression

of harvested power is derived explicitly and validated numerically for the

SSHI system. It is shown that the electrical response using an ideal SSHI

interface is similar to that using the standard interface in a strongly coupled

electromechanical system operated at short circuit resonance. On the other

hand, if the SSHI circuit is not ideal, the performance degradation is

evaluated and classified according to the relative strength of coupling. It is

found that the best use of the SSHI harvesting circuit is for systems in the

mid-range of electromechanical coupling. The degradation in harvested

power due to the non-perfect voltage inversion is not pronounced in this case,

and a new finding shows that the reduction in power is much less sensitive to

frequency deviations than that using the standard technique.

(Some figures in this article are in colour only in the electronic version)

1. Introduction

The need to power remote systems or embedded devices inde-pendently, coupled with advancements in low-power electron-ics, has motivated many research efforts to focus on producing electrical energy from various ambient energy sources. These include solar power, thermal gradients and vibration. Among these energy scavenging sources, parasitic mechanical vibra-tion is a potential power source that is abundant enough to be of use, is easily accessible through microelectromechanical systems (MEMS) manufacturing processes for conversion to electrical energy, and is ubiquitous in applications from small

household appliances to large infrastructure elements [50,52].

While there are several options for transmitting vibra-tion energy into electrical energy, vibravibra-tion-based piezoelec-tric converters have received much attention as

transduc-ers [18,31], since they have high electromechanical coupling,

3 Author to whom any correspondence should be addressed.

require no external voltage source, and are particularly

at-tractive for use in MEMS [10,17, 22, 25, 38]. As a result,

piezoelectric elements for scavenging energy from ambient vibration sources have been used in various types of struc-tures to serve specific purposes, including the use of reso-nant piezoelectric-based structures of cantilever beam

configu-ration [1,11,14,23,40,51,54] as well as plate (membrane)

configuration [7,8,16,29,30,66]. Other harvesting schemes

include the use of long strips of piezoelectric polymers in ocean

or river-water flows [2,61], the use of piezoelectric ‘cymbal’

transducers [27,28], and the use of piezoelectric windmills for

generating electrical power from wind energy [45,46]. The

application to structural health monitoring with vibration

pow-ered piezoelectric sensors can be found in [15,37,41].

Fundamental research on the study of optimal AC power output has been reported extensively in the recent

literature [12, 13, 26, 44, 60, 65]. In most of the work

reported above, the vibration source is typically represented by a single harmonic signal, and an energy scavenger is required

(3)

to resonate mechanically at a frequency tuned to that of the driving source to which it is attached, in order to generate

the maximum electrical energy. As a frequency mismatch

of only a few percent may result in a significant drop in power output, methods to adjust the resonance frequency to match the driving vibrations across a range of frequencies have

been proposed by [6, 34, 53]. Fundamental investigations

concerning the conversion efficiency by piezoelectric power

generators include the early works by [19, 62, 63] and the

recent works by [9, 48,49, 56]. Another fundamental study

relating energy harvesting to electrically induced damping can

be found in [36]. It is shown that the damping added to a

vibrating structure is attributed to the removal of electrical energy from the system, and an explicit expression of this induced damping is provided in the case of a weakly coupled electromechanical system. In addition, the power harvesting system is demonstrated to work similarly to a shunt damping

system [35] except that the energy is stored instead of

dissipated.

The research works cited above focus mainly on developing optimal energy harvesting structures so only AC

power output is considered there. However, the electrical

outputs of these devices in many cases are too small to power electrical devices directly. Thus, the methods of accumulating and storing parasitic energy for intermittent use are also the

key to developing self-powered systems [58,59]. In addition,

a vibrating piezoelectric element generates an AC voltage, while the electronic device in many practical applications requires a stabilized DC voltage. As a result, the electrical interface connecting the piezoelectric element and the terminal electrical load is needed to ensure electrical compatibility. Power optimization schemes therefore depend not only on the mechanical solicitation, but also on the specific type of electronic interface circuit chosen in the energy harvesting

system. However, much work has addressed optimal AC

power flow, while little has considered AC/DC power output

until recently. Ottman et al [42, 43] have used the linear

load impedance adaptation to develop highly efficient electrical

circuits to store the generated charge. They have claimed

that at high levels of excitation the power output can be increased by as much as 400%. But the linear assumption implies that the internal current source is independent of the

load impedance [32]. Shu and Lien [55] have proposed

an improved analysis without this assumption to study the optimal AC/DC power generation for a rectified piezoelectric

device. An accurate estimation of AC/DC power output

is proposed explicitly and is related to electrically induced

damping and energy conversion efficiency [56]. It is shown that

the optimization criteria vary according to the relative strength of the electromechanical coupling. Other related studies can be found in [20,24].

Recently, Guyomar et al [21] have developed a new

electronic interface circuit to enhance power extraction. It is called synchronized switch harvesting on inductor (SSHI). This approach was derived from the so-called synchronized switch

damping (SSD), a nonlinear technique developed originally

by Richard et al [47] to investigate the effect of vibration

damping on mechanical structures. It is shown that the

electrical harvested power can be increased by as much

as 400–900% over the standard technique [4, 5, 21, 39].

However, the approach they used to evaluate the performance of the SSHI circuit is based on the assumption that the periodic excitation and the speed of mass are in-phase. As a result, the effect of frequency deviation from resonance on the electrical behavior of an SSHI system is not included

in their analysis. As vibration-based energy scavenging

generators achieve the maximum power by requiring the matching of resonance frequency to applied frequency, power reduction due to frequency deviation cannot be ignored. Therefore, this paper revisits performance evaluation of the

SSHI electronic interface based on the improved analysis [57].

A new estimate of average harvested power for the SSHI system is derived explicitly and validated numerically, and the connection between the standard and SSHI techniques is also established here. The electrical performance of the SSHI system is evaluated according to two key requirements: (1) the degradation in harvested power due to the non-perfect voltage inversion, and (2) the sensitivity in power deduction

due to frequency deviation. The analysis reveals that the

best use of the SSHI circuit is for systems in the mid-range of electromechanical coupling, since the performance degradations are the least in these cases.

2. Harvesting circuit: a standard interface

Consider an energy conversion device including a vibrating piezoelectric structure together with an energy storage

system. Suppose the modal density of such a device is

widely separated and the structure is vibrating at around its resonance frequency; we may model the power generator as a

mass+spring+damper+piezo structure [21, 42,48]. The

governing equations of this electromechanical model system can be described by

M¨u(t) + ηm˙u(t) + K u(t) + Vp(t) = F(t), (1)

− ˙u(t) + Cp˙Vp(t) = −I (t), (2)

where u(t) is the displacement of the mass M, Vp(t) the

voltage across the piezoelectric element, ηm the mechanical

damping coefficient,K the effective stiffness,the effective

piezoelectric coefficient,Cpthe clamped capacitance,I(t)the

current flowing into the specified circuit andF(t)the external

forcing function. These effective constantsM, K, andCp

are dependent on the properties of the chosen materials and the specific types of piezoelectric generators. For example,

we refer to [13] and [64] for the explicit expressions of

these effective constants for the piezoelectric bimorph power

generator. This paper considers a harmonic mechanical

excitation given by

F(t) = F0sinwt, (3)

where F0 is the constant magnitude and w (in radians per

second) is the angular frequency of vibration. This case is particularly chosen since many applications of piezoelectric materials for power generation involve the use of periodic straining of piezoelectric elements.

The unknowns in (1) and (2) are u(t), Vp(t) and I(t),

while there are only two equations. An additional equation is required and can be obtained from the consideration of the

(4)

Piezoelectric Vibrator Piezoelectric Vibrator

Ce R V (t)c Ce R V (t)c

I(t) I(t)

L

(a) (b)

Figure 1. (a) A standard energy harvesting circuit. (b) An SSHI energy harvesting circuit.

t t ti tf V (t)p Vc -Vc u(t) u0 -u0 T 2 ti tf t t ti tf V (t)p Vc -Vc u(t) u0 -u0 T 2 ti tf VcqI -VcqI (a) (b)

Figure 2. Typical waveforms of displacement and piezoelectric voltage for (a) the standard and for (b) the SSHI electronic interfaces.

chosen electronic interface between the piezoelectric element and the terminal electric load. Since the electronic load needs a stabilized DC voltage while a vibrating piezoelectric element generates an AC voltage, this requires a suitable AC-to-DC

converter to ensure electrical compatibility. For example,

figure1(a) is a standard interface circuit commonly used for

design analysis [42]. It provides an estimation of an upper

bound of the real power that the piezoelectric power generator is able to deliver at a given excitation. The standard interface includes an AC/DC rectifier followed by a filtering capacitance

Ce added to smooth the DC rectified voltageVc. The terminal

load is modeled using an equivalent resistor R with power

assumption equal to the average input power of the actual

terminal electric load. Typically, the filter capacitor Ce is

chosen to be large enough so that the rectified voltage Vc is

essentially constant to have a stable DC output voltage [42]. As

a result, the rectified voltageVcis independent ofCeprovided

that the time constant RCe is much larger than the oscillating

period of the generator [21].

To understand how the standard AC/DC electronic

interface shown in figure 1(a) works for energy transfer, a

perfect rectifying bridge is chosen for demonstration. It is

open circuited if the piezoelectric voltage|Vp| < Vc, and the

outgoing piezoelectric element currentI in (2) is null. On the

other hand, when |Vp| reaches Vc, the bridge conducts and

the piezoelectric voltage is blocked at the rectified voltage;

i.e., |Vp| = Vc. The conduction in the rectifier diodes

is blocked again when |Vp| starts decreasing. Hence, the

piezoelectric voltageVp(t)either varies proportionally with the

displacementu(t)when the rectifying bridge is blocking, or

is kept equal to Vc when the bridge conducts. As the model

equations (1)–(3) are developed at the resonance mode of the

device, a single-mode vibration of the structure is expected.

Thus, the displacement at the steady-state operation is assumed to be

u(t) = u0sin(wt − θ), (4)

where u0 is the magnitude and θ is the phase shift. This

assumption of choosing the sinusoidal form for displacement

has been made by Guyomar et al [21] excluding the effect of

the phase shiftθ. Shu and Lien [55] have included this effect

and validated it both numerically and experimentally for the

standard interface. The corresponding waveforms ofu(t)and

Vp(t)are illustrated in figure2(a).

To analyze the steady-state response of (1) and (2)

connected to the standard AC/DC interface under the harmonic

excitation (3), first note that it can be shown that the rectified

voltage Vc is related to the magnitude of displacement u0

by [21,42,55]

Vc= 2Rw

2CpRw + π

u0. (5)

This formulation (5) can be derived by integrating (2) over a

semi-period of vibration, and by using the fact that the average

current flowing through the capacitanceCe is null at

steady-state operation. Thus, from (5), u0 has to be determined to

decideVc. There are three approaches in the recent literature

for estimating it [21, 42, 55]. The first one models the

piezoelectric device as the current source in parallel with its

internal electrode capacitanceCp [25,41,42]. It is based on

the assumption that the internal current source of the generator

is independent of the external load impedance. However,

the amplitude of the current source is closely related to that of displacement, which depends not only on the mechanical damping but also on the electrical damping at the resonant

vibration [36, 56]. This assumption is therefore not suitable

(5)

result, Guyomar et al [21] have proposed another estimation accounting for the effect of electromechanical coupling. Their estimation is based on the assumption that the external forcing function and the velocity of the mass are in-phase, or in other

words the phase shift effect is neglected in (4). Instead, Shu

and Lien [55] have included this phase factor in their improved

analysis, and derived the analytic expressions of displacement

magnitudeu0, rectified voltageVcand harvested average power

P. Their results are summarized as follows:

u0= u0 F0 K =  1 2ζm+ 2k2er (r+π 2)2 2 2+1− 2+ ke2r r+π 2 212 , (6) Vc= Vc F0  =  r r +π2  × ke2  2ζm+ 2ke2r (r+π 2)2 2 2+  1− 2+ k2er r+π2 212 , (7) P= P F2 0 wscM =  1 r +π 2 2 × ke22r  2ζm+ 2ke2r (r+π 2)2 2 2+  1− 2+ k2er r+π 2 2, (8)

where several non-dimensionless variables are introduced by

k2e =  2 K Cp , ζm= ηm 2√K M, wsc= K M,  = w wsc, r= CpwscR. (9) Above, k2

e is the alternative electromechanical coupling

coefficient,ζmthe mechanical damping ratio,wscthe natural

oscillation frequency (of the piezoelectric vibrator under the

short circuit condition) andandrthe normalized frequency

and electric resistance. Notice that there are two resonances for the system since the piezoelectric structure exhibits both short circuit and open circuit stiffness. They are defined by

sc=1, oc =

1+ k2

e, (10)

wherescandocare the frequency ratios of the short circuit

and open circuit, respectively. Note that the shift in device

natural frequency is pronounced if the coupling factor k2

e is

large.

The improved estimates (6)–(8) have been found to

agree well with experimental observations and numerical

simulations of (1) and (2) [55]. Therefore, these estimates

are suitable for the electrical performance evaluation of the piezoelectric energy harvesting system embedded with the standard electronic interface. Moreover, if the vibration source

is due to the periodic excitation of some base, this givesF0=

M A, whereAis the magnitude of acceleration of the exciting

base. From (8), the harvested average power per unit mass

becomes P M = A2 wscP(r, , ke2, ζm). 0 0.3 0.6 0.9 1.2 1.5 0.8 1 1.2 1.4 1.6 0 0.4 0.8 1.2 1.6 Electromechanical Coupling Factor Normalized Frequency Normalized Power

Figure 3. The normalized power P against the normalized frequency

 and the electromechanical coupling factor k2

eat the optimal conditions in the sense that Popt(, k2

e, ζm) = P(ropt(), , ke2, ζm)

and ropt() is determined by solving

∂rP(r, , ke2, ζm) = 0. We use

ζm= 0.04 for the whole simulation. Notice that for large ke2there are

two identical peaks of power evaluated at the frequency ratio close to

sc= 1 and oc=

1+ k2

e. These peaks are saturated for much higher coupling factor k2

e 1.

Thus, the optimization scheme is closely related to the tuning of the electric resistance, the selection of suitable operation points, and the magnitudes of the coupling coefficient

and mechanical damping ratio. Basically, from (8) the

harvested average power increases significantly for smaller

mechanical damping ratio ζm or larger electromechanical

coupling coefficient k2

e. It is consistent with that found

by Badel et al [3], who have performed an interesting

experiment by comparing the performances of vibration-based piezoelectric power generators using a piezoelectric ceramic and a single crystal. Under the same operating conditions, the power generated using the single crystal is much higher than that using the ceramic, since according to their measurements

the coupling factork2

e of the former is 20 times larger than

that of the latter. However, one has to be cautious that the average harvested power approaches its saturation value for

much largerk2

e, as illustrated in figure3.

The improved estimates have also been compared with the un-coupled and in-phase estimates according to the relative magnitudes of electromechanical coupling coefficient and

mechanical damping ratio. The results of [55] show that

the conventional un-coupled solution and in-phase estimate

are suitable, provided that the ratio ke2

ζm  1, while

the discrepancies among these distinct approaches become

significant when k2e

ζm increases. If the shift in device natural

frequency is pronounced and the mechanical damping ratio of the system is small, i.e. k2e

ζm 1, the harvested power is shown

to have two optima evaluated at(r1opt, opt1 ) and(r2opt, opt2 ),

where opt1 is close to sc and the electric loadr1opt is very

small, while opt2 is close to oc and r2opt is large. Indeed,

table1summarizes the relation between the system parameters

k2

e and ζm and the normalized load, displacement, voltage

and power at these two optimal conditions. The first optimal

pair is designed at the short circuit resonance sc with the

optimal loadrscopt ∝ k12e

ζm

(6)

10–4 10–2 100 102 104 0.8 1 1.2 1.4 1.6 Normalized Frequency Normalized Resistance Normalized Power 0 0.4 0.8 1.2 1.6 10–4 10–2 100 102 104 0.8 1 1.2 1.4 1.6 Normalized Frequency Normalized Resistance Normalized Power 0 0.4 0.8 1.2 1.6 (a) (b)

Figure 4. The normalized power against the normalized electric resistance and frequency ratio. (a) A strongly coupled electromechanical

system using the standard AC/DC electronic interface (k2

e = 1.0, ζm= 0.04,k

2 e

ζm = 25). (b) A weakly coupled electromechanical system using

the ideal SSHI electronic interface (k2

e= 0.01, ζm= 0.04,k

2 e

ζm = 0.25, QI= ∞). Notice that (a) and (b) provide identical peaks of harvested

power evaluated at different conditions.

Table 1. The relation between the system parameters k2

eandζmand the normalized electric resistance, displacement, voltage and power operated at the short circuit (sc) and open circuit (oc)

resonances [55]. Note that the conditionk2e

ζm  1 is implied in the analysis. Optimal conditions sc oc Resistance ropt sc ∝ 1 ke2 ζm < ropt oc ∝ 1 (1+k2 e) k2 e ζm Displacement uopt0 ∝ 1 ζm > u opt 0 ∝ 1 ζm(√1+k2 e) Voltage Voptc ∝ 1 < V opt c ∝√1 1+k2 e k2e ζm Power Popt∝ 1 ζm = P opt ∝ 1 ζm

the open circuit resonanceoc with the optimal loadrocopt ∝

1 (1+k2

e) k2

e

ζm. They give identical values of maximum harvested

power, which depends only on the mechanical damping ratio

ζm. Unlike the power, the displacement is higher atscthan at

oc, while the voltage operating at the first peak is one order

of magnitude smaller than that operating at the second peak.

Finally, figure4(a) gives the dependence of the normalized

harvested power on the normalized resistance and frequency ratio for the case of strong electromechanical coupling. While such a strong coupling is not commonly observed in the conventional piezoelectric power generators, we particularly emphasize it here since we will show in the next section that the behavior of an ideal SSHI system is similar to that of a strongly coupled electromechanical standard system excited at around the short circuit resonance. This finding is generally valid no matter whether the real electromechanical system is weakly or strongly coupled.

3. Harvesting circuit: an SSHI interface

An SSHI electronic interface consists of adding a switch and

an inductanceLconnected in series and is in parallel with the

piezoelectric element as shown in figure1(b). The electronic

switch is triggered according to the maximum and minimum of the displacement of the mass, causing the processing of piezoelectric voltage to be synchronized with the extreme values of displacement.

To illustrate the electrical behavior of this nonlinear processing circuit, consider the harmonic excitation given

by (3). In view of the single-mode excitation, the mechanical

displacement u(t) is assumed to be sinusoidal as in (4) in

steady-state operation. The validation of this assumption will

be examined in section4.1by considering the output voltage.

The waveform of the piezoelectric voltageVp(t), however, may

not be sinusoidal and is dependent on the specific type of the interface circuit connected to the piezoelectric element. To

see this, letT = 2π

w be the period of mechanical excitation

andti andtf be two time instants such that the displacement

u(t)undergoes from the minimum−u0 to the maximumu0

as illustrated in figure2(b). The switch is turned off most of

the time during this semi-period(ti+, tf). When it is turned on

at the time instantti, |Vp(t)|remains lower than the rectified

voltage Vc. So the rectifying bridge is open circuited, and

an oscillating electrical circuit composed by the inductance

L and the piezoelectric capacitanceCp is established, giving

rise to an inversion process for the piezoelectric voltageVp.

Specifically, let tbe the half electric period of this oscillating

LCpcircuit. It is equal to [21]

t = π LCp.

We assume that the inversion process is quasi-instantaneous in the sense that the inversion time is chosen to be much

smaller than the period of the mechanical vibration; i.e., t =

ti+− ti  T. The switch is kept closed during this small

time period t, resulting in the reverse of voltage on the

piezoelectric element, i.e.,

Vp(ti+) = −Vp(ti)e

−π

2QI = VcqI, qI=e

−π

(7)

as illustrated in figure2(b). Above,QIis the inversion quality factor due to the energy loss mainly from the inductor in series with the switch. As a result, the current outgoing from the piezoelectric element through the rectifier during a half

vibration period can be obtained by integrating (2) from time

ti+totf [3] tf ti+  − ˙u(t) + Cp˙Vp(t)  dt= −2u0 + Cp  1−e−2QπI  Vc = −T 2 Vc R

since the rectifier bridge is blocking during the inversion

process and the inversion time t  T. The relation between

the magnitude of displacementu0and the rectified voltageVc

is therefore obtained by

Vc= 2Rw

(1− qI)CpRw + π

u0. (12)

The rest of the problem is to estimate the magnitude of

displacementu0and the phase shiftθ. To solve them, consider

the balance of energy first. Let (1) be multiplied by˙u(t)and (2)

be multiplied byVp(t). Integration of the addition of these two

equations from timeti+totf gives the equation of the energy

balance tf ti+ F(t) ˙u(t)dt = tf ti+ ηm˙u2(t)dt+ tf ti+ Vp(t)I (t)dt +1 2M˙u 2(t)|tf ti++ 1 2K u 2(t)|tf ti++ 1 2CpV 2 p(t)| tf ti+ = tf ti+ ηm˙u2(t)dt+ tf ti+ Vp(t)I (t)dt+12Cp(1− qI2)V 2 c (13) where tf ti+ F(t) ˙u(t)dt= π 2F0u0sinθ, tf ti+ ηm˙u2(t)dt= π 2ηmwu 2 0, tf ti+ Vp(t)I (t)dt = π w V2 c R.

There are now three unknowns,u0, θ andVc, while only

two equations, (12) and (13), are found. The third equation

connecting these unknowns can be obtained as follows. Notice that from (2)

 ˙Vp(t) = 

Cp[−I (t) +  ˙u(t)].

(14)

Differentiating (1) with respect to timetand substituting (14)

into it, we find

M d dt¨u(t) + ηm d dt˙u(t) +  K + 2 Cp  d dtu(t) −  Cp I(t) = d dtF(t). (15)

Integrating (15) with respect to time t from ti+ to tf and

using (4) provides the third equation

 K− Mw2+ 2 Cp  u0π 2CpwR Vc= F0cosθ. (16)

Thus, the unknownθ can be eliminated from (13) and (16).

This gives F02=  ηmwu0+  2 wR+ (1− qI2) Cp π  V2 c u0 2 +  K− Mw2+ 2 Cp  u0− π 2CpwR Vc 2 . (17)

The above equation (17) can be further simplified to findu0by

using the relation betweenu0and Vc given by (12), and this

gives u0= F0 ⎧ ⎨ ⎩  ηm+2  1+C p Rw2π (1−qI2)  2R ( 1−qI) 2 CpwR+π2 2 2 w2+  K−w2M+ ( 1−qI) 2 w2R (1−qI) 2 CpwR+π2 2 ⎫ ⎬ ⎭ 1 2 .

The average harvested power is therefore obtained onceu0is

determined since from (12)

P= V 2 c R = 4R2w2 {(1− qI)CpRw + π}2 u20. (18) The following summarizes our main results for the piezoelectric element connected to the SSHI interface circuit.

The normalized displacement magnitude uSSHI

0 , rectified

voltage VSSHIc and average harvested power PSSHI are given

respectively by uSSHI0 =u SSHI 0 F0 K = 1 ⎧ ⎨ ⎩  2ζm+2[1+ r 2π(1−qI2)]k2er (1−q I) 2 r+ π 2 2 2 2+  1−2+ ( 1−qI) 2 k2er (1−qI) 2 r+π2 2 ⎫ ⎬ ⎭ 1 2 , (19) VSSHIc = V SSHI c F0  =  r (1−qI) 2 r + π2  × ke2 ⎧ ⎨ ⎩  2ζm+2[1+ r 2π(1−q2I)]k2er (1−q I) 2 r+π2 2 2 2+  1−2+ ( 1−qI) 2 k2er (1−qI) 2 r+ π 2 2 ⎫ ⎬ ⎭ 1 2 , (20) PSSHI= P SSHI F2 0 wscM =  1 (1−qI) 2 r +π2 2 × ke22r  2ζm+2[1+ r 2π(1−q2I)]ke2r ( 1−qI) 2 r+ π 2 2 2 2+  1− 2+ ( 1−qI) 2 k 2 er (1−qI) 2 r+π2 2. (21) All above are expressed in terms of non-dimensionless

parameters defined in (9).

Finally, Guyomar et al [21] have used the in-phase

assumption to analyze the electrical performance of the power generator using the SSHI interface. To be precise, they have assumed that the external forcing function and the velocity of

(8)

Normalized Frequency Normalized Power r : 0.13 r : 0.28 r : 0.48 r : 5.5 r : 10 r : 20 r : 0.82 r : /2 r : 3 Normalized Frequency r : 0.13 r : 0.28 r : 0.48 r : 0.82 r :π/2 r : 3 r : 5.5 r : 10 r : 20 N ormalize dP o w e r (a) (b) 0 0.3 0.6 0.9 1.2 1.5 0.9 0.95 1 1.05 1.1 1.15 0 0.3 0.6 0.9 1.2 1.5 1.1 1.15 0.9 0.95 1 1.05 π

Figure 5. Normalized harvested power versus frequency ratio for various load resistances based on (a) the in-phase estimate PSSHIin−phaseand

(b) the improved analytic estimate PSSHI. We take k2

e= 0.09, ζm= 0.04 and QI= ∞ for the whole simulation.

the mass are in-phase, giving rise to no phase shift effect in their formulation. The following summarizes their results for future comparisons: uSSHIin−phase=u SSHI in−phase F0 K =  1 2ζm+2[1+ r 2π(1−qI2)]k2er (1−q I) 2 r+π2 2   , (22) VSSHIin−phase= V SSHI in−phase F0  =  r (1−qI) 2 r +π2  × ke2 2ζm+2[1+ r 2π(1−q2I)]ke2r (1−q I) 2 r+π2 2 , (23) PSSHIin−phase= P SSHI in−phase F2 0 wscM = 1 (1−qI) 2 r +π2 2 × k2er  2ζm+2[1+ r 2π(1−q2I)]ke2r ( 1−qI) 2 r+ π 2 2 2. (24)

4. Discussion

4.1. In-phase versus improved analysis

The in-phase estimate shows a lack of frequency dependence. To see this, we consider a case where the ratio of the electromechanical coupling factor to the mechanical damping

ratio is in the medium range; i.e., k2e

ζm is of order one. We

take k2

e = 0.09 andζm = 0.04, and therefore

k2 e

ζm = 2.5. It

can be shown that other parameters provide similar contrasting comparisons between the in-phase and improved estimates.

Figure 5(a) describes the average harvested power versus

frequency ratio for various load resistances based on the

in-phase estimate while figure 5(b) is based on our improved

analysis. For fair comparison, the ideal inversion of voltage

is implied in this case; i.e., QI = ∞. We will discuss

this condition in more detail later. Clearly, the two estimates predict identical values of the optimal power. However, the in-phase estimate is unable to predict the system behavior when the applied driving frequency deviates from the system resonance frequency. As the reduction in power is significant due to frequency deviation, such an effect cannot be ignored in practical design.

We next validate our improved estimates numerically by

transforming (1)–(3) to an equivalent circuit with R∗ = ηm

2

as resistance, L∗ = M2 as inductance and C∗ = 

2 K as

capacitance as shown in figure 6(a). We use the software

PSpice to simulate this equivalent circuit connected to the

SSHI interface. We takeke2 =0.09, ζm= 0.04 and consider

the non-ideal voltage inversion with quality factor QI =

2.6 [21]. The results are illustrated in figure 6(b), where

the normalized power is plotted against the frequency ratio

evaluated at the optimal electric load,ropt =1.01 in this case.

The predicted results from the in-phase and analytic estimates

are represented by dashed and solid lines in figure6(b), while

the numerical results are marked using open circles there. Apparently, the numerical simulation favors results predicted based on our analytic improved estimate. Therefore, from

now on, we will use (19)–(21) as the fundamental tool for the

performance evaluation of the piezoelectric energy harvesting system embedded with an SSHI interface circuit.

4.2. Ideal inversion of piezoelectric voltage

To see how the SSHI electronic interface boosts power extraction, consider an ideal case where the inversion of the

piezoelectric voltageVpis complete; i.e.,QI= ∞. From (11)

(9)

Ce Vc I(t) Cp L Piezoelectric Vibrator R R* L* C* (a) 0.9 0.95 1 1.05 1.11 1.15 Normalized Frequency 0 0.4 0.8 1.2 1.6 Normalized P ower Simulation Analytic Estimate In-Phase Estimate (b)

Figure 6. Numerical validation with parameters k2

e= 0.09, ζm= 0.04 and QI= 2.6. (a) The equivalent circuit model for the system of

equations (1)–(3) connected to the SSHI interface. (b) The simulation results compared with those predicted by analytic and in-phase estimates. becomes PSSHI= 4 π2 r k2 e2  4  ζm+4k2 er π2 2 2+ (1− 2)2 . (25)

The optimal electric load resistance and the normalized power

operated atscare therefore

ropt= π 2 4 1 k2 e ζm , PSSHI|r=ropt,=1= 1 16ζm. (26)

From (26), the optimal load resistance is inversely proportional

to the ratio k2e

ζm while the corresponding optimal power depends

only on the mechanical damping ratioζmand is independent of

the electromechanical coupling coefficientk2

e. Comparing all

of these features with table1suggests that the behavior of the

power harvesting system using the SSHI interface is similar to that of a strongly coupled electromechanical system using the standard interface and operated at the short circuit resonance

sc. Indeed, comparing (5) with (12) suggests that the original

capacitanceCpis replaced by the effective capacitance ¯Cp =

(1−qI)

2 Cp. Therefore, the effective electromechanical coupling

coefficient¯ke2can be realized by

¯k2 e = 2 K ¯Cp = 22 K(1− qI)Cp . (27)

For the ideal inversion,qI → 1, and this gives ¯ke2

ζm → ∞no

matter what the original value of the ratio k2e

ζm is. The response

using the SSHI interface is therefore similar to that using the standard interface in a strongly coupled electromechanical

system operated atsc. In addition, according to table1, there

exists another identical peak of power operated at the open circuit resonance. But now this frequency ratio is realized as

¯oc =

1+ ¯k2

e → ∞. Hence, the second peak of power is

moved to the infinite point in the(r, )space, and therefore

there is only one peak of power for the SSHI electronic interface, no matter whether the real electromechanical system is weakly or strongly coupled, as schematically shown in figure4(b).

In addition, we particularly takek2

e =0.01 andζm=0.04

in figure4(b) so that the electromechanical generator itself is

weakly coupled(k2e

ζm = 0.25). The harvested power obtained

using the standard harvesting circuit is pretty small in this case,

since it has been shown that [55]

P  ropt= π 2,  =1, k 2 e, ζm  ≈  2 π ke2 ζm  1 16ζm  1 16ζm =  PSSHI  max if k2e

ζm  1. But the inclusion of the SSHI circuit boosts

the average harvested power, whose maximum is the same as that using a strongly coupled electromechanical generator

(10)

(k2

e =1.0,ζm=0.04 and

k2 e

ζm =25). Therefore, the harvested

power increases tremendously for any weak coupling SSHI system at the cost of using a much larger optimal electric load,

which is proportional to 1

ke2

ζm

according to (26).

5. Comparisons

We now compare the electrical performance of a vibration-based piezoelectric power generator using the standard and SSHI electronic interfaces according to the different ratios of ke2

ζm. As in many practical situations, the inversion of the

piezoelectric voltage Vp is not perfect (QI = ∞), which

accounts for a certain amount of the performance degradation

using the SSHI electronic interface. We take QI = 2.6

for comparison here [21]. It is possible to have a larger

value of quality factor QI by requiring the use of a low loss

inductor [33].

First, consider a weakly coupled electromechanical

system; i.e., the ratio k2e

ζm  1. We take k

2

e = 0.01

and ζm = 0.04 for demonstration. This gives k2e

ζm =

0.25. The harvested power versus frequency ratio for various

normalized resistances is shown in figure 7(a) based on the

standard interface and in figure 7(d) based on the SSHI

interface. The maximum normalized power generated for the

ideal voltage inversion is around PSSHI|QI=∞ = 1.56, while

PSSHI|QI=2.6 = 0.67 in the non-ideal case. In spite of the significant performance degradation in this case, the achieved optimal power is three times larger than that using the standard

interface (P=0.23) at the cost of using a larger matching load

resistance fromropt = π/2 to 5.2 by comparing figure7(a)

with figure7(d).

Next, suppose the electromechanical coupling is in the

medium range; i.e., the ratio of k2e

ζm is of order one. We take

k2

e = 0.09 and ζm = 0.04. This gives

k2 e

ζm = 2.25. The

harvested power versus frequency ratio for various normalized

resistances is shown in figure 7(b) based on the standard

interface and in figure 7(e) based on the SSHI interface.

The maximum normalized power for the non-ideal voltage

inversion is PSSHI|QI=2.6 = 1.38, which is slightly smaller

than the ideal case (PSSHI|QI=∞ = 1.56), but it is slightly

larger than that using the standard electronic interface (P =

1.20). While there is no significant increase of power output

using the SSHI electronic interface in this case, figure 7(e)

demonstrates that the harvested power evaluated at around the optimal load is less sensitive to frequency deviating from the resonant vibration. For example, the amount of normalized

harvested power P evaluated atr = π2 in the standard case

drops from 1.2 to 0.6 for about 5% frequency deviation, and from 1.2 to 0.2 for about 10% frequency deviation. However, under the same conditions, the normalized harvested power

PSSHI in the SSHI circuit drops from 1.3 to only 1.0 for

about 5% frequency deviation, and from 1.3 to 0.5 for about 10% frequency deviation. This frequency-insensitive feature is much more pronounced in the case of ideal voltage inversion,

as can be seen by comparing figure5(b) with figure7(b).

Finally, we turn to a strongly coupled electromechanical system(ke2

ζm 1). Our numerous numerical simulations show

that the rule of thumb for the appearance of double identical peaks of power in the standard case is when

k2 e ζm 10.

We then take k2

e = 1.0 and ζm = 0.04, and this gives

k2 e

ζm = 25. The harvested power versus frequency ratio for

various normalized resistances is shown in figure7(c) based

on the standard interface and in figure7(f) based on the SSHI

interface. In the standard case, the harvested power has two identical optimal peaks, and the switching between these two

peaks can be achieved by varying the electric loads. The

envelope of these peaks has a local minimum, which is closely related to the minimum proof mass displacement. On the other hand, there is only one peak of power in the SSHI circuit, as explained in the previous section. Unlike the standard case

as illustrated in figure7(c), the peaks of the average harvested

power decrease significantly as the load resistances increase, as

shown in figure7(f). In addition, it can be seen from (26) that

the optimal electric load for the SSHI system is very small, since ke2

ζm  1. Thus, figure7(f) indicates that any deviation

in the load resistance will cause a significant power drop in the SSHI case. Such an effect cannot be ignored in practical design, since there may exist other inherent electrical damping in the whole circuit system; for example, the diode loss is not taken into account in the present analysis. As a result, there seems to be no obvious advantage in using the SSHI electronic

interface from the comparison between figures7(c) and (f).

6. Conclusions

The electrical behavior of the piezoelectric power harvesting system using the SSHI electronic interface is analyzed and

compared to that using the standard interface. Instead of

using the un-coupled or in-phase assumptions to estimate the harvested power, an analytic expression of it based on the improved analysis is proposed explicitly and validated numerically for the SSHI circuit system. It is found that no matter whether the real electromechanical system is weakly or strongly coupled, the electrical response using an ideal SSHI interface is similar to that using the standard interface in a strongly coupled electromechanical system operated at the short circuit resonance. As a result, the harvested power increases tremendously for any weak coupling SSHI system at the cost of using a much larger optimal electric load.

The performance degradation due to the non-perfect voltage inversion is discussed and classified according to the relative strength of the coupling. This effect on power deduction is significant for weakly coupled electromechanical

systems. On the other hand, if the electromechanical

coupling is in the medium range, the degradation in harvested power is not pronounced in this case, and a new finding shows that the reduction in power is much less sensitive to frequency deviations than that using the standard technique. This provides a great advantage in design since the energy scavenger has a wider inherent bandwidth. Moreover, this frequency-insensitive feature is much more conspicuous when the inversion quality factor is improved.

(11)

0.9 Normalized Frequency 0.95 1 1.05 1.1 0 0.15 0.3 0.45 0.6 0.75 Normalize d P ower r : 0.4 r : 0.52 r : 0.7 r : 0.1 r : /2 r : 2.8 r : 5.2 r : 8 r : 12 0.9 Normalized Frequency 0.95 1 1.05 1.1 0 0.15 0.3 0.45 0.6 0.75 Normalized P ower r : 0.4 r : 0.52 r : 0.7 r : 0.1 r : /2 r : 2.8 r : 5.2 r : 8 r : 12 (a) (d) 0.9 Normalized Frequency 0.95 1 1.05 1.1 1.15 0 0.3 0.6 0.9 1.2 1.5 Normalize dP ower r : 0.13 r : 0.28 r : 0.45 r : 0.75 r : /2 r : 3 r : 5.5 r : 10 r : 20 r : 0.13 r : 0.28 r : 0.45 r : 0.75 r : /2 r : 3 r : 5.5 r : 10 r : 20 0.9 Normalized Frequency 0.95 1 1.05 1.11 1 .15 0 0.3 0.6 0.9 1.2 1.5 Normaliz ed Power (b) (e) r : 0.1 r : 0.25 r : 0.42 r : 0.7 r : 1.2 r : 2 r : 3.2 r : 5.3 r : 13 0.8 1 1.2 1.4 1.6 Normalized Frequency Normalized Powe r 0 0.25 0.5 0.75 1 1.25 1.5 1.75 r : 0.1 r : 0.25 r : 0.42 r : 0.7 r : 1.2 r : 2 r : 3.2 r : 5.3 r : 13 0.8 1 1.2 1.4 1.6 Normalized Frequency Norma lized Power 0 0.25 0.5 0.75 1 1.25 1.5 1.75 (c) (f ) π π π π

Figure 7. Normalized power versus frequency ratio for different values of normalized resistances. Notice that (a)–(c) are obtained using the

(12)

Acknowledgments

We thank Professors K C Wu, C K Lee and U Lei for

their constant encouragement and support. We are glad

to acknowledge partial support from the National Science Council under grant No 96-2628-E-002-119-MY3, and from the Ministry of Economic Affairs under grant No 95-EC-17-A-05-S1-017 (WHAM-BioS).

References

[1] Ajitsaria1 J, Choe S Y, Shen D and Kim D J 2007 Modeling and analysis of a bimorph piezoelectric cantilever beam for voltage generation Smart Mater. Struct.16 447–54

[2] Allen J J and Smits A J 2001 Energy harvesting EEL J. Fluids

Struct.15 629–40

[3] Badel A, Benayad A, Lefeuvre E, Lebrun L, Richard C and Guyomar D 2006 Single crystals and nonlinear process for outstanding vibration-powered electrical generators IEEE

Trans. Ultrason. Ferroelectr. Freq. Control 53 673–84

[4] Badel A, Guyomar D, Lefeuvre E and Richard C 2005 Efficiency enhancement of a piezoelectric energy harvesting device in pulsed operation by synchronous charge inversion

J. Intell. Mater. Syst. Struct.16 889–901

[5] Badel A, Guyomar D, Lefeuvre E and Richard C 2006 Piezoelectric energy harvesting using a synchronized switch technique J. Intell. Mater. Syst. Struct.17 831–9

[6] Charnegie D, Mo C, Frederick A A and Clark W W 2006 Tunable piezoelectric cantilever for energy harvesting Proc.

2006 ASME Int. Mechanical Engineering Congr. and Exposition IMECE2006–14431

[7] Cho J, Anderson M, Richards R, Bahr D and Richards C 2005 Optimization of electromechanical coupling for a thin-film PZT membrane: I. Modeling J. Micromech. Microeng.

15 1797–803

[8] Cho J, Anderson M, Richards R, Bahr D and Richards C 2005 Optimization of electromechanical coupling for a thin-film PZT membrane: II. Experiment J. Micromech. Microeng.

15 1804–9

[9] Cho J H, Richards R F, Bahr D F, Richards C D and Anderson M J 2006 Efficiency of energy conversion by piezoelectrics Appl. Phys. Lett.89 104107

[10] Choi W J, Jeon Y, Jeong J H, Sood R and Kim S G 2006 Energy harvesting MEMS device based on thin film piezoelectric cantilevers J. Electroceram.17 543–8

[11] Cornwell P J, Goethal J, Kowko J and Damianakis M 2005 Enhancing power harvesting using a tuned auxiliary structure J. Intell. Mater. Syst. Struct.16 825–34

[12] duToit N E and Wardle B L 2007 Experimental verification of models for microfabricated piezoelectric vibration energy harvesters AIAA J.45 1126–37

[13] duToit N E, Wardle B L and Kim S G 2005 Design considerations for MEMS-scale piezoelectric mechanical vibration energy harvesters Integrated Ferroelectr.

71 121–60

[14] Elvin N G, Elvin A A and Spector M 2001 A self-powered mechanical strain energy sensor Smart Mater. Struct.

10 293–9

[15] Elvin N G, Lajnef N and Elvin A A 2006 Feasibility of structural monitoring with vibration powered sensors Smart

Mater. Struct.15 977–86

[16] Ericka M, Vasic D, Costa F, Poulin G and Tliba S 2005 Energy harvesting from vibration using a piezoelectric membrane

J. Physique IV 128 187–93

[17] Fang H B, Liu J Q, Xu Z Y, Dong L, Chen D, Cai B C and Liu Y 2006 A MEMS-based piezoelectric power generator for low frequency vibration energy harvesting Chin. Phys.

Lett.23 732–4

[18] Flatau A B and Chong K P 2002 Dynamic smart material and structural systems Eng. Struct.24 261–70

[19] Goldfarb M and Jones L D 1999 On the efficiency of electric power generation with piezoelectric ceramic Trans. ASME,

J. Dynam. Syst. Meas. Control 121566–71 [20] Guan M J and Liao W H 2007 On the efficiencies of

piezoelectric energy harvesting circuits towards storage device voltages Smart Mater. Struct.16 498–505

[21] Guyomar D, Badel A, Lefeuvre E and Richard C 2005 Toward energy harvesting using active materials and conversion improvement by nonlinear processing IEEE Trans. Ultrason.

Ferroelectr. Freq. Control52 584–95

[22] Horowitz S B, Sheplak M, Cattafesta L N III and Nishida T 2006 A MEMS acoustic energy harvester

J. Micromech. Microeng.16 S174–81

[23] Hu H P, Cao J G and Cui Z J 2007 Performance of a piezoelectric bimorph harvester with variable width

J. Mech. 23 197–202

[24] Hu H P, Xue H and Hu Y T 2007 A spiral-shaped harvester with an improved harvesting element and an adaptive storage circuit IEEE Trans. Ultrason. Ferroelectr. Freq. Control

54 1177–87

[25] Jeon Y B, Sood R, Jeong J H and Kim S G 2005 MEMS power generator with transverse mode thin film PZT Sensors

Actuators A122 16–22

[26] Jiang S, Li X, Guo S, Hu Y, Yang J and Jiang Q 2005 Performance of a piezoelectric bimorph for scavenging vibration energy Smart Mater. Struct.14 769–74

[27] Kim H W, Batra A, Priya S, Uchino K, Markley D, Newnham R E and Hofmann H F 2004 Energy harvesting using a piezoelectric ‘cymbal’ transducer in dynamic environment Japan. J. Appl. Phys.43 6178–83

[28] Kim H W, Priya S, Uchino K and Newnham R E 2005 Piezoelectric energy harvesting under high pre-stressed cyclic vibrations J. Electroceram.15 27–34

[29] Kim S, Clark W W and Wang Q M 2005 Piezoelectric energy harvesting with a clamped circular plate: analysis J. Intell.

Mater. Syst. Struct.16 847–54

[30] Kim S, Clark W W and Wang Q M 2005 Piezoelectric energy harvesting with a clamped circular plate: experimental study

J. Intell. Mater. Syst. Struct.16 855–63

[31] Lee C K, Hsu Y H, Hsiao W H and Wu J W J 2004 Electrical and mechanical field interactions of piezoelectric systems: foundation of smart structures-based piezoelectric sensors and actuators, and free-fall sensors Smart Mater. Struct.

13 1090–109

[32] Lefeuvre E, Badel A, Richard C and Guyomar D 2005 Piezoelectric energy harvesting device optimization by synchronous electric charge extraction J. Intell. Mater. Syst.

Struct.16 865–76

[33] Lefeuvre E, Badel A, Richard C, Petit L and Guyomar D 2006 A comparison between several vibration-powered

piezoelectric generators for standalone systems Sensors

Actuators A126 405–16

[34] Leland E S and Wright P K 2006 Resonance tuning of piezoelectric vibration energy scavenging generators using compressive axial preload Smart Mater. Struct.15 1413–20

[35] Lesieutre G A 1998 Vibration damping and control using shunted piezoelectric materials Shock Vib. Dig.30 187–95

[36] Lesieutre G A, Ottman G K and Hofmann H F 2004 Damping as a result of piezoelectric energy harvesting J. Sound Vib.

269 991–1001

[37] Liao W H, Wang D H and Huang S L 2001 Wireless monitoring of cable tension of cable-stayed bridges using PVDF piezoelectric films J. Intell. Mater. Syst. Struct. 12 331–9 [38] Lu F, Lee H P and Lim S P 2004 Modeling and analysis of

micro piezoelectric power generators for

micro-electro-mechanical-systems applications Smart Mater.

Struct.13 57–63

[39] Makihara K, Onoda J and Miyakawa T 2006 Low energy dissipation electric circuit for energy harvesting Smart

(13)

[40] Mateu L and Moll F 2005 Optimum piezoelectric bending beam structures for energy harvesting using shoe inserts

J. Intell. Mater. Syst. Struct.16 835–45

[41] Ng T H and Liao W H 2005 Sensitivity analysis and energy harvesting for a self-powered piezoelectric sensor J. Intell.

Mater. Syst. Struct.16 785–97

[42] Ottman G K, Hofmann H F, Bhatt A C and Lesieutre G A 2002 Adaptive piezoelectric energy harvesting circuit for wireless remote power supply IEEE Trans. Power Electron.

17 669–76

[43] Ottman G K, Hofmann H F and Lesieutre G A 2003 Optimized piezoelectric energy harvesting circuit using step-down converter in discontinuous conduction mode IEEE Trans.

Power Electron.18 696–703

[44] Poulin G, Sarraute E and Costa F 2004 Generation of electric energy for portable devices: comparative study of an electromagnetic and a piezoelectric system Sensors

Actuators A116 461–71

[45] Priya S 2005 Modeling of electric energy harvesting using piezoelectric windmill Appl. Phys. Lett.87 184101

[46] Priya S, Chen C T, Fye D and Zahnd J 2005 Piezoelectric windmill: a novel solution to remote sensing Japan. J. Appl.

Phys.44 L104–7

[47] Richard C, Guyomar D, Audigier D and Ching G 1998 Semi passive damping using continuous switching of a piezoelectric device Proc. SPIE3672 104–11

[48] Richards C D, Anderson M J, Bahr D F and Richards R F 2004 Efficiency of energy conversion for devices containing a piezoelectric component J. Micromech. Microeng.

14 717–21

[49] Roundy S 2005 On the effectiveness of vibration-based energy harvesting J. Intell. Mater. Syst. Struct.16 809–23

[50] Roundy S, Leland E S, Baker J, Carleton E, Reilly E, Lai E, Otis B, Rabaey J M, Wright P K and Sundararajan V 2005 Improving power output for vibration-based energy scavengers IEEE Pervasive Comput.4 28–36

[51] Roundy S and Wright P K 2004 A piezoelectric vibration based generator for wireless electronics Smart Mater. Struct.

13 1131–42

[52] Roundy S, Wright P K and Rabaey J M 2004 Energy

Scavenging for Wireless Sensor Networks with Special Focus on Vibrations (Boston, MA: Kluwer–Academic)

[53] Shahruz S M 2006 Design of mechanical band-pass filters with large frequency bands for energy scavenging Mechatronics

16 523–31

[54] Shenck N S and Paradiso J A 2001 Energy scavenging with shoe-mounted piezoelectrics IEEE Micro21 30–42

[55] Shu Y C and Lien I C 2006 Analysis of power output for piezoelectric energy harvesting systems Smart Mater. Struct.

15 1499–512

[56] Shu Y C and Lien I C 2006 Efficiency of energy conversion for a piezoelectric power harvesting system J. Micromech.

Microeng.16 2429–38

[57] Shu Y C and Lien I C 2007 A comparison between the standard and SSHI interfaces used in piezoelectric power harvesting

Proc. SPIE: Active and Passive Smart Struct. Integr. Syst.

6525 652509

[58] Sodano H A, Inman D J and Park G 2005 Generation and storage of electricity from power harvesting devices J. Intell.

Mater. Syst. Struct.16 67–75

[59] Sodano H A, Lloyd J and Inman D J 2006 An experimental comparison between several active composite actuators for power generation Smart Mater. Struct.15 1211–6

[60] Stephen N G 2006 On energy harvesting from ambient vibration J. Sound Vib. A293 409–25

[61] Taylor G W, Burns J R, Kammann S M, Powers W B and Welsh T R 2001 The energy harvesting Eel: a small subsurface ocean/river power generator IEEE J. Oceanic

Eng.26 539–47

[62] Umeda M, Nakamura K and Ueha S 1996 Analysis of the transformation of mechanical impact energy to electric energy using piezoelectric vibrator Japan. J. Appl. Phys.

35 3267–73

[63] Umeda M, Nakamura K and Ueha S 1997 Energy storage characteristics of a piezo-generator using impact induced vibration Japan. J. Appl. Phys.36 3146–51

[64] Wang Q M and Cross L E 1999 Constitutive equations of symmetrical triple layer piezoelectric benders IEEE Trans.

Ultrason. Ferroelectr. Freq. Control46 1343–51

[65] Williams C B and Yates R B 1996 Analysis of a micro-electric generator for microsystems Sensors Actuators A52 8–11

[66] Yang J, Chen Z and Hu Y T 2007 An exact analysis of a rectangular plate piezoelectric generator IEEE Trans.

數據

Figure 2. Typical waveforms of displacement and piezoelectric voltage for (a) the standard and for (b) the SSHI electronic interfaces.
Figure 3. The normalized power P against the normalized frequency
Figure 4. The normalized power against the normalized electric resistance and frequency ratio
Figure 5. Normalized harvested power versus frequency ratio for various load resistances based on (a) the in-phase estimate P SSHI in−phase and (b) the improved analytic estimate P SSHI
+3

參考文獻

相關文件

6 《中論·觀因緣品》,《佛藏要籍選刊》第 9 冊,上海古籍出版社 1994 年版,第 1

Wang, Unique continuation for the elasticity sys- tem and a counterexample for second order elliptic systems, Harmonic Analysis, Partial Differential Equations, Complex Analysis,

Based on [BL], by checking the strong pseudoconvexity and the transmission conditions in a neighborhood of a fixed point at the interface, we can derive a Car- leman estimate for

Robinson Crusoe is an Englishman from the 1) t_______ of York in the seventeenth century, the youngest son of a merchant of German origin. This trip is financially successful,

fostering independent application of reading strategies Strategy 7: Provide opportunities for students to track, reflect on, and share their learning progress (destination). •

Strategy 3: Offer descriptive feedback during the learning process (enabling strategy). Where the

It should be stressed that the four eigenvalues obtained here do not change even if we include other field outside KBc subalgebra or outside the dressed B 0 gauge, since such fields

In this work, for a locally optimal solution to the nonlin- ear SOCP (4), under Robinson’s constraint qualification, we show that the strong second-order sufficient condition