You are on page 1of 13

Model Formulation

The model of the single-degree-of-freedom (SDOF) VI-EH system is a nonlinear differential equation
(NDE) and the equation of motion of the moving levitated magnet inside the dual-purpose device is
given as

mz c1 z k1 z k3 z 3 my (1)

Z Y
where z x y, x X sin( t ) , y Y sin( t ) , z Z sin( t ) and Ta then the Vibration
Y
Isolator Energy Harvester (VI-EH) model becomes,

mz c1 z k1 z k3 z 3 m 2Y sin( t ) (2)

The dynamic model of the VI-EH system as described in (2) is representative of the duffing-equation.
Such equations have previously been solved using several mathematical tools like the method of
multiple scales [1], direct numerical integration [2] , nonlinear normal forms (NLNF) [3] and
harmonic balance method [4],[5]. In this study, the output frequency response function (OFRF) will
be employed for the analysis, design and optimisation of the VI-EH system.
The average input mechanical power of the VI-EH system due to the base acceleration is given as

1
Pin mY 2 3
(3)

However the average output power (power harvested) across the load resistance of the harvesting
circuit is given as
2
1 kt Z
Pout RL (4)
2 Rc RL

where kt BNl

Considering the average output power, Pout of the VI-EH system when subjected to a harmonic base
excitation with an average input mechanical power of Pin , the energy conversion efficiency  e of the
VI-EH system is given as;
Pout
e  %    100 (5)
Pin

The output response of system (3), Z is a function of the excitation frequency, and the nonlinear
parameter, k 3 . Therefore considering (4) and (5), it can be deduced that the average output power, Pout
and energy conversion efficiency,  e are both functions of these parameters. Though the VI-EH
system has a dual-purpose function, its primary function is vibration isolation where the absolute
displacement transmissibility is less than unity i.e. Ta  1 while the secondary function is energy
harvesting. This implies the frequency range of interest is n 2 which comprises the region of

isolation hence energy harvesting can only happen within this region. The objective here is to
maximise e within the isolation range Ta  1 . The next section discusses the OFRF concept and
thereafter the OFRF polynomials of the relative displacement transmissibility, absolute displacement

1
transmissibility and subsequently the average output power and energy conversion efficiency is
derived.

The OFRF Concept

The OFRF of a nonlinear system is derived from a given nonlinear differential equation model
wherein a polynomial relationship between the output frequency response of the system and the
parameters defining the system nonlinearities is derived [6]. The OFRF provides an analytical
representation of the output frequency response of the system in terms of the systems nonlinear
parameters and thus describes the characteristics of the system. This concept is valid for the class of
nonlinear systems stable at zero equilibrium. An extensive study of the OFRF concept can be found in
[6]–[13]. It should be noted, in this study, that the following terms - OFRF
‘Polynomial/approximation/representation’ is used interchangeably.

Consider the Volterra systems described by the differential equation


M m L p
d li z (t ) m
d li y (t )
c p ,m p (l1 , lm ) 0 (6)
m 1 p 0 l1 ,...lm 0 i 1 dt li i p 1 dt
li

where L is the order of the derivative and M is the maximum degree of nonlinearity in terms of the
system input and output, y(t ) and z (t ) . The system output response of (2) can be represented by a
polynomial function in terms of the system parameters as
m1 mSN
j1 jSN
Z(j ) ( j1 , , jSN ) (j ) 1 SN (7)
j1 0 jSN 0

where Z ( j ) represents the output spectrum of (6), ( j1 , , jSN ) ( j ) are complex-valued frequency
functions (also called ‘OFRF coefficients’) of the system linear parameters and system input while
j1 jSN
1 SN is a set of monomials (OFRF structure) in the OFRF representation of the output spectrum.
Let the set of monomials in the OFRF representation of the nth-order output spectrum of (3) be
denoted as M and the frequency function vector be denoted as ( j ) , therefore the OFRF
polynomial is obtained as

Z(j ) M ( j )T (8)

where

(j ) 0 , 1, 2 , 3, 4 , 5, 6 , 7 and
N (9)
M Mn
n 1

Determination of the OFRF structure (monomials)


In [8], [12], an algorithm was proposed used to determine the structure of the OFRF which is
employed herein to determine M n thus

2
L n 1 n (m p) L
Mn c0, n (l1 , , ln ) c p ,( m- p ) (l1 , , lm ) Mn ( m p ), p
l1 , ,ln 0 m p 1 p 1 l1 , ,ln 0
(10)
n L
c p ,0 (l1 , , lm ) Mn , p
p 2 l1 , ,ln 0

where the character ‘  ’ is the Kronecker product and


n p 1
M n, p Mi Mn i, p 1 , M n,1 M n , M1 [1] (11)
i 1

N
Then the set of monomials is obtained as M Mn
n 1

System Analysis
In this section, the OFRF method is used in the analytical study of system (2). System (2) is a
particular instance of (6) for L  2 and M  3 with system parameters obtained as c10 (2)  m ,
c10 (1)  c1 , c10 (0)  k1 , c30 (000)  k3 , c01 (0) m 2Y , else cp ,m p 0 . Applying the algorithm for
obtaining the OFRF structure (monomials) as presented in (10) and (11) to system (2) up to 7th-order
i.e. N=15 , yields the following monomials;
N
M Mn 1, k3 , k32 , k33 , k34 , k35 , k36 , k37 (12)
n 1

Determination of the frequency function vector

The frequency function vector ( j ) is computed by applying the method described in proposition 4
outlined in [6]. Fifteen simulation studies were conducted where system (2) was excited with same
base input but the nonlinear stiffness parameter, k 3 takes different values from a training set
k3  0 : 0.1:1.4  106 . The output spectra of the system are evaluated from the 15 sets of numerically
simulated (Runge-Kutta method) system outputs to obtain the results
Z ( j ) |k (i ) for i 1, 2,3......,15 and the estimates of the OFRF coefficients are evaluated using the least
3

square method as

Z(j ) k
0(j ) 3 (1)

Z(j ) k
1( j ) 3 (2)
T 1 T
(j ) (13)
6(j ) Z(j ) k
3 (14)

7(j ) Z(j ) k
3 (15)

where

3
1 k3 (1) k32 (1) k36 (1) k37 (1)
1 k3 (2) k32 (2) k36 (2) k37 (2)
(14)

1 k3 (15) k32 (15) k36 (15) k37 (15)

Therefore, the OFRF of system (2), when subjected to a specific input, is obtained as

Z ( j , k3 ) 0 (j ) 1 ( j ) k3 2 ( j )k32 3 ( j )k33 6 ( j )k36 7 ( j )k37


R (15)
Z ( j , k3 ) r ( j )k3r where R 7
r 0

where ( j ) are the frequency functions dependent on the system input and linear characteristic
parameters, c1 and k1 .

Using the algorithm proposed in [12] , the square magnitude of the output response of (15) can be
computed as
R R
2
Z ( j , k3 ) Z ( j , k3 ) Z ( j , k 3 ) r ( j )k3r r ( j ) k3r
r 0 r 0
t
2 *
Z ( j , k3 ) 0 0 k3t *
t
t 1 0 (16)
2 2 13 14
Z ( j , k3 ) 0 1 3 k k
2 3 13 3k k
14 3
R
2
Z ( j , k3 ) r ( )k3r where R 14
r 0

Eq.(16d) can be substituted in (4) thus;


2
1 kt RL
Pout ( , k3 ) Z2
2 RL Rc RL
2
R
1 kt RL
r ( )k3r (17)
2 RL Rc RL r 0
2
R
r 1 kt RL
( ) r ( )k 3 where ( )
r 0 2 RL Rc RL

Similarly, substituting (3) and (17) in (5) yields


2
R
1 kt RL
r ( )k3r
2 RL Rc RL r 0
e ( , k3 )
1 3 2
m Y
2
R
kt RL
2 r ( )k3r (18)
2mRL 2
Y Rc RL r 0

2
R
r kt RL
( ) r ( )k 3 where ( )= 2
r 0 2mRL 2
Y Rc RL

4
From (17) and (18), the following representations for Pout and e can be made.
R
Pout ( , k3 ) r ( ) k3r where r ( ) ( ) r ( )
r 0
R
(19)
r
e ( , k3 ) r ( )k 3 where r ( ) ( ) r ( )
r 0

Eq. (19) shows the OFRF approximations for the average output power, Pout and the energy
conversion efficiency, e of the VI-EH system respectively. The OFRF for the absolute displacement
transmissibility, Ta , can also be derived as in (13) thus
R
Ta ( , k3 ) r ( ) k3r (20)
r 0

where r ( ) , r ( ) and r ( ) are frequency functions dependent on the system input and linear
characteristic parameters.

System design and optimisation

In designing the natural frequency and the linear damping coefficient of the system at
n 55.6 rad.s-1 and c1  2.7 Nsm-1 respectively, two design equality constraints can be established
thus;

 f1 (c1 , k1 , k3 ) : k1  24.87  0
 (21)
 f 2 (c1 , k1 , k3 ) : c1  2.7  0

The OFRF polynomials for the system performance indices of interest Ta , Pout ,e  are obtained
subject to the bound constraint k3  1.4 106 Nm-3 . Given the derived OFRF polynomial
approximations, the optimisation problem can be formulated accordingly;
max e (c1 , k1 , k3 )
c1 , k1 , k3

f1 (c1 , k1 , k3 ) : k1 24.87 0
f 2 (c1 , k1 , k3 ) : c1 2.7 0 (22)
s.t. f3 (c1 , k1 , k3 ) : k3 1.4 10 6
0
R
f 4 (c1 , k1 , k3 ) : r ( ) k3r 1 0
r 0

In this study, four levels of base acceleration inputs, 0.25g, 0.5g, 0.75g and 1g , are considered. For
each input, the maximum  e obtainable by the VI-EH system, subject to the set of constraints on the
system, is computed using a MATLAB optimisation tool ‘fmincon’ and the results presented in table
1.

Results and discussions

Prior to using the OFRF models derived for the analysis, design or optimisation of the system under
consideration, there has to be a form of validation. This is to confirm that the OFRF models derived
for a specific system input, given the system parameters, describes the actual system dynamics over

5
the entire spectrum considered. In Fig. (2), this validation is presented for all base excitation levels
considered, however for a value of k3  1.5 106 Nm-3 which is outside the OFRF training (design)
range of k3  0, 1.4  106 Nm-3 . It is observed that the OFRF approximations evidently matches the
ODE45 numerical simulations for all system inputs considered (0.25g, 0.5g, 0.75g and 1g). This
certainly establishes the efficiency of the OFRF method. In Table 1, the maximum energy conversion
efficiency achievable with the corresponding k 3 values, subject to system constraints and base
acceleration levels, is summarised.
Table 1: Maximum energy conversion efficiency attainable at specific nonlinear stiffness value subject to
the system constraints at several base excitations
Base accel. Absolute displacement Max. Energy conversion Nonlinear stiffness,
level transmissibility , Ta efficiency, emax k3 (Nm-3 )

0.25g 0.9869 0.7624 1.4  106


0.50g 0.9914 0.7659 1.4  106
0.75g 0.9996 0.7723 1.4  106
1.00g 0.9999 0.7724 0.8  106

It is observed that the energy conversion efficiency increased with base acceleration level within the
constraint set, Ta  1 and k3  0, 1.4  106 Nm-3 . It is also seen that k3  1.4 106 Nm-3 is required to
achieve emax for base accelerations o.25g, 0.5g and 0.75g due to the constraint on k 3 however, for
base acceleration of 1g, k3  0.8 106 Nm-3 is required to obtain emax due to the constraint on Ta .

In designing the VI-EH system over the isolation region ( n 2 ), it is apparent, as presented in

Fig. (3a) that at low levels of base acceleration, Ta increases with c1 while at high level excitation, it
decreases with c1 . For a linear system  k3  0  , over the isolation region and for all base excitation
levels, absolute displacement transmissibility increases with increased linear damping (limitation of
linear damping in vibration isolation). However, the system considered in this study is nonlinear as a
result of the hardening stiffness,  k3  0  term and this typical nonlinearity causes a right-shift in the
resonant frequency of the system. The effect of the nonlinearity becomes distinct as the excitation
level increases therefore, at high excitation levels, a shift in the resonant frequency results to a shift in
the isolation region.

A close inspection of Fig. (3b) shows that while the energy conversion efficiency,  e is sensitive to
linear damping, c1 , it is independent of the excitation level. This is possibly so since for any linear
damping value, the average output power increases as the average input mechanical power increases
 Pout 
 e  %    100  thus keeping the efficiency constant.
 Pin 

The energy conversion efficiency,  e may be independent of the excitation level but the average
output power is definitely not as depicted in Fig. (3c). It is seen that though the average output power
decreases with an increasing linear damping, it increases as the excitation level increases for specific
damping levels. It is also apparent that the rate of decrease of Pout as c1 increases, is greater for high
level excitations compared to the low levels.

In Fig. (4a), there is a clear trend of an increase in Ta as k 3 increases, nonetheless, the gradient of the
line is large (steep) for high excitation levels and small (flat) for low excitation levels. This further
reinforces the effect of different excitation levels on nonlinearities. At k3  0 , the system is linear and

6
Ta is independent of the excitation level. Similarly, as illustrated in Fig. (4b),  e is independent of the
excitation level at k3  0 though it increases, at a rate dependent on the excitation level, as k 3
increases.

Unlike Ta and  e , even at high excitation levels, Pout is barely sensitive to k 3 as demonstrated in Fig.
(4c). However, Pout is very sensitive to the excitation level and shows a significant increase as
excitation level increases.

The effect of varying the load resistance, RL while keeping the coil resistance, Rc constant, is also
considered in Figs. (5a) and (5b). In Fig. (5a), maximum  e is obtained when RL  Rc for all
excitation levels. This further confirms the conclusions made from Fig. (3b) as RL creates a linear
electromagnetic damping effect in the mechanical subsystem of the VI-EH system. Nevertheless, in
Fig. (5b), the effect of varying the excitation levels is clearly seen as Pout increases with the base
excitation level and also, like in Fig. (5a), the maximum Pout is achieved when RL  Rc .

7
(a) Numerical (ODE45) vs OFRF solution at Base excitation of 0.25g

(b) Numerical (ODE45) vs OFRF solution at Base excitation of 0.5g

(c) Numerical (ODE45) vs OFRF solution at Base excitation of 0.75g

(d) Numerical (ODE45) vs OFRF solution at Base excitation of 1g

8
Figure 2: (a)-(d) Comparison between ODE45 numerical simulation to the corresponding OFRF for
different base acceleration levels (0.25g, 0.5g, 0.75g and 1g)

9
10
11
12
References
[1] T. D. Equation and N. Oscillators, The Duffing Equation. .
[2] A. Erturk and D. J. Inman, “Broadband piezoelectric power generation on high-energy orbits
of the bistable Duffing oscillator with electromechanical coupling,” J. Sound Vib., vol. 330,
no. 10, pp. 2339–2353, 2011.
[3] A. Cammarano, S. Neild, S. Burrow, D. Wagg, and D. Inman, “Optimum resistive loads for
vibration-based electromagnetic energy harvesters with a stiffening nonlinearity,” J. Intell.
Mater. Syst. Struct., vol. 25, no. 14, pp. 1757–1770, 2014.
[4] S. Souayeh and N. Kacem, “Computational models for large amplitude nonlinear vibrations of
electrostatically actuated carbon nanotube-based mass sensors,” Sensors Actuators, A Phys.,
vol. 208, pp. 10–20, 2014.
[5] A. Jallouli, N. Kacem, G. Bourbon, P. Le Moal, V. Walter, and J. Lardies, “Pull-in instability
tuning in imperfect nonlinear circular microplates under electrostatic actuation,” Phys. Lett. A,
vol. 380, no. 46, pp. 3886–3890, 2016.
[6] Z. Q. Lang, S. a. Billings, R. Yue, and J. Li, “Output frequency response function of nonlinear
Volterra systems,” Automatica, vol. 43, pp. 805–816, 2007.
[7] Z.-Q. Lang and S. . Billings, “Output frequencies of a class of nonlinear systems,” Int. J.
Control, vol. 67, no. 5, pp. 713–730, 1997.
[8] Z.-Q. LANG and S. A. BILLINGS, “Output frequency characteristics of nonlinear systems,”
Int. J. Control, vol. 64, no. 6, pp. 1049–1067, Feb. 2007.
[9] X. J. Jing, Z. Q. Lang, and S. A. Billings, “Output frequency properties of nonlinear systems,”
Int. J. Non. Linear. Mech., vol. 45, no. 7, pp. 681–690, 2010.
[10] J. C. Peyton Jones and K. Choudhary, “Output frequency response characteristics of nonlinear
systems. Part II: overlapping effects and commensurate multi-tone excitations,” Int. J. Control,
vol. 85, no. 9, pp. 1279–1292, 2012.
[11] Z. Q. Lang, P. F. Guo, and I. Takewaki, “Output frequency response function based design of
additional nonlinear viscous dampers for vibration control of multi-degree-of-freedom
systems,” J. Sound Vib., vol. 332, no. 19, pp. 4461–4481, 2013.
[12] X. J. Jing, Z. Q. Lang, and S. A. Billings, “Output frequency response function-based analysis
for nonlinear Volterra systems,” Mech. Syst. Signal Process., vol. 22, no. 1, pp. 102–120,
2008.
[13] Y. Zhu and Z. Q. Lang, “Design of Nonlinear Systems in the Frequency Domain: An Output
Frequency Response Function-Based Approach,” IEEE Trans. Control Syst. Technol., pp. 1–
14, 2017.

13

You might also like