You are on page 1of 12

Extremely Large Oscillations

of Cantilevers Subject
to Motion Constraints
Hamed Farokhi The nonlinear extremely large-amplitude oscillation of a cantilever subject to motion
Department of Mechanical and constraints is examined for the first time. In order to be able to model the large-
Construction Engineering, amplitude oscillations accurately, the equation governing the cantilever centerline rota-

Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020


Northumbria University, tion is derived. This allows for analyzing motions of very large amplitude even when tip
Newcastle Upon Tyne NE1 8ST, UK angle is larger than p/2. The Euler–Bernoulli beam theory is employed along with the
e-mail: hamed.farokhi@northumbria.ac.uk centerline inextensibility assumption, which results in nonlinear inertial terms in the
equation of motion. The motion constraint is modeled as a spring with a large stiffness
Mergen H. Ghayesh1 coefficient. The presence of a gap between the motion constraint and the cantilever
School of Mechanical Engineering, causes major difficulties in modeling and numerical simulations, and results in a non-
University of Adelaide, smooth resonance response. The final form of the equation of motion is discretized via the
Adelaide 5005, South Australia, Australia Galerkin technique, while keeping the trigonometric functions intact to ensure accurate
e-mail: mergen.ghayesh@adelaide.edu.au results even at large-amplitude oscillations. Numerical simulations are conducted via a
continuation technique, examining the effect of various system parameters. It is shown
that the presence of the motion constraints widens the resonance frequency band effec-
tively which is particularly important for energy harvesting applications.
[DOI: 10.1115/1.4041964]

Keywords: cantilever, motion constraint, extremely large oscillation, nonlinear damping,


bistability

1 Introduction stability of a bar subjected to a periodic driving force, a linear


damping force, and impact forces due to its motion.
Temporary contact and impact between components are present
Most of the recent studies on this topic focus on the application
in many mechanical structures and machines, such as piping sys-
of motion constraints in enhancing the efficiency of vibration-
tems, vibration isolators, gears with clearances, and impact damp-
based energy harvesters via widening the resonance frequency
ing elements. Impact between components is also present in a
band. For instance, Liu et al. [15] used a lumped model to analyze
specific class of energy harvesters [1,2]. For all these systems,
the response of a piezoelectric energy harvester subject to motion
accurate modeling and analysis of the dynamical behavior of the
constraints. Soliman et al. [16] developed an energy harvester
system is essential for predicting the motion amplitudes as well as
design implementing motion constraints to widen the frequency
the developed force and stress magnitudes, to ensure smooth oper-
band. Similar theoretical and experimental investigations have
ation and prevent improper functioning and immature failure of
been performed on design and analysis of vibration-based energy
system components.
harvesters by, for instance, Halim and Park [17], Wu et al. [18],
The early studies on this topic focused on investigating the
and Liu et al. [1]. All of these valuable studies are based on exper-
behavior of piecewise linear discrete systems [3–10]. For instance,
imental observations and/or discrete (lumped mass) theoretical
Moon and Shaw [3] studied the chaotic vibrations of an elastic
analysis.
beam with nonlinear boundary conditions. Nigm and Shabana
The present study, for the first time, investigates the nonlinear
[11] conducted a theoretical analysis on the steady-state vibra-
extremely large-amplitude oscillation of a cantilever under base
tional characteristics of a discrete system subject to an impact
excitation subject to motion-limiting constraints from both sides,
damper. Further investigation was conducted by Shaw and
employing a continuous model. In particular, a continuous model
Holmes [4], who studied vibrational behavior of a single degree-
of the cantilever is developed on the basis of the Euler–Bernoulli
of-freedom (DOF) piecewise linear oscillator. Choi and Noah [5]
beam theory and inextensibility condition. The centerline rotation
examined the forced response of a single DOF system involving
is considered as the main variable describing the motion of the
an unsymmetrical piecewise linear stiffness using a harmonic bal-
cantilever; this allows analyzing very large-amplitude motions
ance technique together with the fast Fourier transformation. The
accurately. The effect of the motion constraint is accounted for
investigations were continued by Natsiavas [9], who examined the
via a spring with a clearance. The presence of an initial gap (clear-
dynamics of oscillators with bilinear damping and stiffness. There
ance) between the cantilever and the motion constraint causes
are also a few studies examining some aspects of the dynamics of
major challenges in continuous modeling and numerical simula-
constrained continuous systems [12,13]. For instance, Metallidis
tions. The continuous model is discretized employing the Galerkin
and Natsiavas [14] examined the axial vibration response of a
technique and solved using a continuation technique. Numerical
periodically excited deformable rod subject to motion-limiting
results are presented in the form of frequency-amplitude dia-
constraint. Bj€orkenstam [12] studied the motion and dynamic
grams, time histories, and phase-plane portraits.

1
Corresponding author. 2 Model Development
Contributed by the Applied Mechanics Division of ASME for publication in the
JOURNAL OF APPLIED MECHANICS. Manuscript received July 22, 2018; final manuscript
The system under consideration, as shown in Fig. 1, is a cantile-
received November 7, 2018; published online December 17, 2018. Assoc. Editor: ver under base excitation subjected to motion constraints. The
George Haller. motion of the cantilever is analyzed in a Cartesian coordinate

Journal of Applied Mechanics Copyright V


C 2019 by ASME MARCH 2019, Vol. 86 / 031001-1
Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020
Fig. 1 (a) Schematic of a cantilever under transverse base excitation subject to motion con-
straints and (b) deformed configuration of the system

system with x and z denoting the axial and transverse directions, @e


respectively. The cantilever is of length L, width b, and thickness Ee þ g
r ¼ |{z} (4)
@t
|{z}
h. The motion constraints are located at a distance xc from the re
rv
base, with a gap width of g0 from each side; both motion con-
straints have the same stiffness. The whole cantilever system is
under harmonic base excitation of amplitude w0 and frequency x. where E and g stand for Young’s modulus and material viscosity,
One of the well-known assumptions in modeling of cantilevers respectively, and re and rv denote the elastic and viscous compo-
is the inextensibility of the centerline. Under this assumption, the nents of the axial stress, respectively.
longitudinal displacement, u(x,t), and the transverse displacement The strain energy of the cantilever is given by [20]
(with respect to the base), w(x,t), can be written as a function of ð ðL  2
the rotation angle of the centerline, h(x,t), as 1 1 @hð x; tÞ
P¼ re e dV ¼ EI dx (5)
ðx 2 V 2 0 @x
uðx; tÞ ¼ ðcos ½hðx; tÞ  1Þdx
where I represents the second moment of area.
ð0x (1) The virtual work done by the viscous component of the axial
wðx; tÞ ¼ sin ½hðx; tÞdx stress can be formulated as
0
ð ðL 2  
@ hð x; tÞ @hð x; tÞ
The total transverse displacement, wt(x,t), is given by dWv ¼  rv de dV ¼ gI d dx (6)
ðx V 0 @x@t @x
wt ðx; tÞ ¼ wðx; tÞ þ w0 sinðx tÞ ¼ sin ½hðx; tÞdx þ w0 sinðx tÞ The kinetic energy of the cantilever under base-excitation can be
0
(2) obtained as

The relations in Eq. (1) allow for deriving the rotational equation ðL  2
1 @hð x; tÞ
of motion of the cantilever. The advantage of the rotational equa- KE ¼ qI dx
2 0 @t
tion of motion over the transverse equation of motion is that it is 
ðL ðx 2
capable of predicting very-large-amplitude oscillations even when 1 @hð x; tÞ  
þ qA sin hð x; tÞ dx dx
the tip angle is larger than p/2. In what follows, the kinetic and 2 0 0 @t
potential energies as well as the work of damping are obtained in ð L ð x 2
1 @hð x; tÞ  
terms of the rotation angle. þ qA cos hð x; tÞ dx þ w0 x cosðx tÞ dx
The axial strain of the cantilever under inextensibility assump- 2 0 0 @t
tion can be formulated as [19] (7)

@hð x; tÞ in which q denotes the mass density and A stands for the cross-
e ¼ z (3) sectional area.
@x
The motion constraints are modeled via linear springs with very
Based on the Kelvin–Voigt energy dissipation mechanism, the large stiffness coefficients. Hence, the work of the motion con-
axial stress can be expressed as straints (both sides) can be mathematically represented as

031001-2 / Vol. 86, MARCH 2019 Transactions of the ASME


ðL ( ðL "  ðx  ! ðx !  ðx  !# )
   
dWc ¼  k1 cos h dD ð x  xc ÞH  sin h dx  g0 sgn sin h dx  sin h dx  g0 dx dh dx (8)
   
0 x 0 0 0

where H denotes the Heaviside function, sgn stands for the sign function, dD represents the Dirac Delta function, and the vertical bars,
i.e., jj, indicate the absolute value.
Employing generalized Hamilton’s principle
ð t2
ðdKE  dP þ dWv þ dWc Þdt ¼ 0 (9)
t1
one can derive the equation governing the rotational motion of the cantilever under transverse base-excitation and subject to motion
constraints as

Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020


" # ( " # )
ðx ðx  2 ðx ð x  2
@2h @2h @3h @h @2h @h @ 2
h
qI 2  EI 2  gI  qAsin h cos h þ 2 sin h dx dx þ qAcosh w0 x2 sinðx tÞ þ sin h  2 cos h dx dx
@t @x @t@x2 @t @t L 0 @t @t
L0
ðL "  ðx  ! ð x   ðx  ! #
   
þk1 cosh dD ð x  xc ÞH  sin hdx  g0 sgn sin h dx  sin hdx  g0
  dx ¼ 0 (10)
x 0 0 0

Introducing the following dimensionless quantities:


x AL2 xc t g w0
x ¼ ; xc ¼ ; t ¼ ; gd ¼ ; w0 ¼
a¼ ; ;
L I L s Es L
g0 k1 L3
xe ¼ xs; g0 ¼ ; K1 ¼ (11)
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi L EI
where s ¼ qAL4 =ðEIÞ, inserting them in Eq. (10) and disregarding the asterisk notation results in
" # ( " # )
ðx ðx  2 ðx ð x  2
1 @2h @2h @3h @h @2h @h @2h
  gd  sin h cos h þ 2 sin h dx dx þ cos h w0 xe sinðxe tÞ þ sin h  2 cos h dx dx
a @t2 @x2 @t@x2 @t @t 1 0 @t @t
10
ð1 "  ðx  ! ð x   ðx  ! #
   
þK1 cos h dD ð x  xc ÞH  sin h dx  g0 sgn sin h dx  sin h dx  g0
  dx ¼ 0
x 0 0 0

(12)

Wiring the equation of motion in dimensionless form allows for a general parametric analysis [21]. In order to be able to examine the
nonlinear dynamical characteristics of the cantilever numerically, Eq. (12) is reduced to a discretized set nonlinear ordinary differential
equations utilizing the Galerkin technique. To this end, the cantilever centerline rotation is expanded as
XN
hðx; tÞ ¼ pm ðtÞwm ðxÞ;
m¼1 (13)
wm ðxÞ ¼ ½sinhðbm xÞ þ sinðbm xÞ  vm ðcoshðbm xÞ  cosðbm xÞÞ
where pm ðtÞ denotes the mth generalized coordinate for the rotational motion and vm ¼ ½coshðbm Þ þ cosðbm Þ=½sinhðbm Þ þsinðbm Þ; bm
is the mth root of the equation 1 þ coshðxÞ cosðxÞ ¼ 0.
Substitution of Eq. (13) into the equation of motion, i.e., Eq. (12), and application of the Galerkin method results in
ð1 ! ð1 ! ð1 !
1 XN X N XN
00 00
w w dx p€n  wm wn dx pn gd wm wn dx p_ n
a n¼1 0 m n n¼1 0 n¼1 0
8 !ð ð 2 !!2 ! ! !3 9
ð1 < XN x x XN XN 2 X N XN =
4 @ @
 wm sin pn ðtÞwn ð xÞ pn ðtÞwn ð xÞ cos pn ðtÞwn ð xÞ þ 2 pn ðtÞwn ð xÞ sin pn ðtÞwn ð xÞ 5dxdx dx
0
: 1 0 @t n¼1 @t n¼1 ;
n¼1 n¼1 n¼1
8 !ð 0 2 ! !2 !
ð1 < X ðx
N x
@ 4 @ X N XN
þ wm cos pn ðtÞwn ð xÞ w0 xe sinðxe tÞ þ pn ðtÞwn ð xÞ sin pn ðtÞwn ð xÞ
0
: 1 0 @t n¼1
n¼1 n¼1
! !#! )
@2 X N X N
 2 pn ðtÞwn ð xÞ cos pn ðtÞwn ð xÞ dxdx dx
@t n¼1 n¼1
8 !ð " !  !
ð1 < X ð x X
N 1  N 
þ wm K1 cos pn ðtÞwn ð xÞ dD ð xxc Þ H  sin pn ðtÞwn ð xÞ dx g0
0 : n¼1 x 0 n¼1

ðx ! ! 0 ð !  !# )
X N  x X N 
sgn sin ð Þw
pn t n ð xÞ dx @  ð Þw 
pn t n ð xÞ dx g0 dx dx ¼ 0; m ¼ 1;2;…;N
 sin
0 n¼1 0 n¼1

(14)

Journal of Applied Mechanics MARCH 2019, Vol. 86 / 031001-3


Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020
Fig. 2 Frequency-amplitude plots of the cantilever without constraint: (a) maximum tip transverse dis-
placement, (b) maximum tip longitudinal displacement, and (c) maximum tip angle. The solid line showing
stable solution, while the dotted line showing unstable one. P1, P2, and P3 are points of interest which are
examined in more detail in Fig. 3.

In order to ensure accurate results even at very large-amplitude cantilever subject to motion constraint is examined and discussed
oscillations, the trigonometric functions of h are kept intact as in Sec. 3.2. The numerical results are obtained for a cantilever of
they are integrated from 0 to 1, which results in substantially large dimensions h ¼ 1 mm, b ¼ 5 h, and L ¼ 200 h. It should be men-
equations of motion. The presence of such large equations and tioned here once again that all the reported values in this section
nonlinear inertial terms makes the numerical simulation even are dimensionless as defined in Eq. (11) with the asterisk notation
more challenging. In this study, five generalized coordinates are being dropped for convenience.
retained in the rotational motion series expansion, resulting in a
5DOF system which ensures converged results; in other words,
3.1 Nonlinear Response of the Cantilever Without Motion
the contribution of higher modes (more than 5) to the cantilever
Constraint. In the absence of a motion constraint, the cantilever
response is small enough that can be safely neglected. After
is free to undergo large-amplitude oscillations. The transverse
obtaining the amplitude of the rotational motion, the longitudinal
amplitude of oscillation depends of course on the base-excitation
and transverse displacements are calculated using Eq. (1). The
amplitude. Figure 2 shows the nonlinear frequency-amplitude
5DOF discretized system of equations is solved numerically via
plots of the cantilever undergoing large-amplitude oscillations due
the use of a pseudo-arclength continuation technique; addition-
to transverse base-excitation. For this case, w0 ¼ 0.018 and
ally, a stability analysis was conducted using the Floquet theory.
gd ¼ 0.0057. Subfigures (a, b) show the cantilever tip transverse
The employed method used is capable of capturing all types of
and longitudinal displacements, while subfigure (c) shows the can-
periodic motions as well as detecting different bifurcations [22].
tilever tip angle. It is worth mentioning that solving the model
developed in Sec. 2 results in the value of centerline angle. Hav-
ing obtained the centerline angle, the transverse and longitudinal
3 Numerical Results displacements are calculated via using Eq. (1). As seen in Fig. 2,
The nonlinear large-amplitude oscillations of the cantilever the system shows a weak nonlinear hardening behavior. Further-
with and without motion constraints are studied in this section. In more, subfigure (c) shows that the tip angle maximum value is
order to better illustrate the capability of the developed model in more than p/2, indicating that the tip of the cantilever bends more
capturing very large-amplitude oscillations of the cantilever, the than 90 deg at large-amplitude oscillations. This is the main
motion constraint is initially removed and the nonlinear resonant advantage of the model developed in this study, as it overcomes
response of the system under base-excitation is examined; this is the limitation of the classical nonlinear transverse model of the
discussed in Sec. 3.1. Then, the detailed nonlinear response of the cantilever (i.e., the tip angle being between p/2 and p/2). To

031001-4 / Vol. 86, MARCH 2019 Transactions of the ASME


xc ¼ 0.2. The frequency-amplitude diagrams are constructed for
different cases and the effect of different parameters is examined
in detail.
Figure 4 shows a comparison between constrained and uncon-
strained motion of the cantilever in the primary resonance region.
For this case, an initial gap width of 0.03 is considered, i.e.,
g0 ¼ 0.03. Furthermore, a large spring stiffness coefficient is con-
sidered, i.e., K1 ¼ 2.0  104, to properly model the effect of the
motion constraint. The base-excitation amplitude is set to
w0 ¼ 0.018 while its frequency (xe) is varied in the vicinity of the
fundamental natural frequency of the cantilever, i.e., x1 ¼ 3.5160.
As seen in subfigure (a), the presence of a motion constraint

Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020


reduces the amplitude of the oscillation, but yields wider reso-
Fig. 3 ((a)–(c)) Oscillation of the system of Fig. 2 at points P1, nance frequency band. This is particularly interesting and useful
P2, and P3, respectively for energy harvesting applications. More specifically, one of the
better illustrate the oscillation of the cantilever, the system motion techniques used in increasing the efficiency of vibration-based
is captured at different frames in one full period for points P1, P2, energy harvesters is to increase the band of resonance vibration,
and P3, and plotted in Fig. 3. As seen in subfigure (a), the cantile- i.e., increase the interval at which the amplitude is larger than a
ver longitudinal displacement is much smaller than its transverse certain value. If a transverse displacement threshold of 0.45 is
displacement; however, at larger oscillation amplitudes, for considered, the cantilever system without constraint undergoes
instance, subfigure (c), the longitudinal displacement becomes oscillations of amplitude larger than 0.45 in the frequency range
comparable to the transverse displacement. of 0.980 < xe/x1 < 1.031; however, the constrained cantilever
system undergoes oscillations of amplitude larger than 0.45 in
the frequency range of 0.980 < xe/x1 < 1.160, i.e., 3.5 times the
3.2 Nonlinear Response of the Cantilever With Motion range of unconstrained system. Comparing the area under the
Constraint. This section examines the nonlinear resonant curve for both constrained and unconstrained systems in the fre-
response of the cantilever in the presence of a motion constraint at quency ratio range shown in the figure (i.e., 0.90 < xe/x1 < 1.18)

Fig. 4 Comparison between the frequency-amplitude plots of the constrained cantilever to those of a
cantilever with no constraint: (a) maximum tip transverse displacement, (b) maximum tip longitudinal dis-
placement, and (c) maximum tip angle. The solid line showing stable solution, while the dotted line show-
ing unstable one. g0 5 0.03, K1 5 2.0 3 104, and w0 5 0.018.

Journal of Applied Mechanics MARCH 2019, Vol. 86 / 031001-5


Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020

Fig. 5 Details of the motion of the system of Fig. 4 at xe/x1 5 0.9775 (i.e., in the resonance region before hitting
the motion constraint), showing, respectively, the time trace and phase-plane portrait of: ((a), (b)) tip transverse
displacement, ((c), (d)) tip longitudinal displacement, and ((e), (f)) tip angle. tn denotes normalized time with
respect to the period of the oscillation.

reveals a value of for the 0.0790 unconstrained system and a value before and after hitting the motion constraint. As seen in Fig. 5,
of 0.1292 for the constrained, i.e., an increase of 63.5%. the cantilever displays smooth periodic response in the region
The detailed dynamical characteristics of the constrained canti- where there is not contact between the cantilever and the con-
lever at two different excitation frequencies, xe/x1 ¼ 0.9775 and straint. However, as seen in Fig. 6, the dynamical characteristics
1.1030, are illustrated in Figs. 5 and 6, respectively. xe/ of the cantilever in the region where it hits the constraint are
x1 ¼ 0.9775 is in the resonance region before hitting the motion totally different, showing nonsmooth vibrational behavior.
constraint, while xe/x1 ¼ 1.1030 is in the region where the canti- It would be interesting to compare the response of the cantile-
lever is hitting the motion constraint. The goal here is to compare ver subject to motion constraint at both sides to that subject to
the time traces and phase-plane portraits of the cantilever motion motion constraint only at one side; the comparison between the

031001-6 / Vol. 86, MARCH 2019 Transactions of the ASME


Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020

Fig. 6 Details of the motion of the system of Fig. 4 at xe/x1 5 1.1030 (i.e., in the resonance region after hitting
the motion constraint), showing, respectively, the time trace and phase-plane portrait of ((a), (b)) tip transverse
displacement, ((c), (d)) tip longitudinal displacement, and ((e), (f)) tip angle. tn denotes normalized time with
respect to the period of the oscillation.

two cases is shown in Fig. 7. As seen, when the cantilever motion important to note that asymmetric motions can arise even in oscil-
is constrained from both sides, the change in the slope of the reso- lators with perfectly symmetric properties.
nance response after hitting the constraint is larger compared to As mentioned in Sec. 2, in this study, the motion constraints are
the case with one-side constraint. Furthermore, the cantilever with modeled as linear springs located at distance xc from the clamped
constraints at both sides shows wider resonance frequency band base and at both sides of the cantilever with a gap (clearance) g0.
compared to the cantilever with a constraint only at one side. It To this end, the spring stiffness coefficient should be large enough
should be noted that for the rest of the figures in this paper, it is to ensure that the spring is a realistic representative of the motion
assumed that the cantilever is constrained at both sides. It is constraint. The effect of the spring stiffness coefficient on the

Journal of Applied Mechanics MARCH 2019, Vol. 86 / 031001-7


Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020
Fig. 7 Comparison between the frequency-amplitude plots of the cantilever constrained at both sides to
those of a cantilever constrained at one side: (a) maximum tip transverse displacement and (b) maximum
tip longitudinal displacement. The solid line showing stable solution, while the dotted line showing unsta-
ble one. g0 5 0.03, K1 5 2.0 3 104, and w0 5 0.018.

Fig. 8 Effect of the motion constraint stiffness on frequency-amplitude plots of the constrained cantile-
ver: (a) maximum tip transverse displacement and (b) maximum tip longitudinal displacement. The solid
line showing stable solution, while the dotted line showing unstable one. g0 5 0.03 and w0 5 0.018.

Fig. 9 Effect of the motion constraint gap-width on the frequency-amplitude plots of the constrained
cantilever: (a) maximum tip transverse displacement and (b) maximum tip longitudinal displacement.
The solid line showing stable solution, while the dotted line showing unstable one. K1 5 2.0 3 104 and
w0 5 0.018.

031001-8 / Vol. 86, MARCH 2019 Transactions of the ASME


Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020
Fig. 10 Effect of base-excitation amplitude on the frequency-amplitude plots of the constrained cantile-
ver: (a) maximum tip transverse displacement and (b) maximum tip longitudinal displacement. The solid
line showing stable solution, while the dotted line showing unstable one. g0 5 0.03 and K1 5 2.0 3 104.

Fig. 11 Effect of the position of the motion constraint on the frequency-amplitude plots of the cantilever:
(a) maximum tip transverse displacement and (b) maximum tip longitudinal displacement. The solid line
showing stable solution, while the dotted line showing unstable one. g0 5 0.03, K1 5 2.0 3 104, and
w0 5 0.018.

frequency-amplitude diagrams of the constrained cantilever is occurs at smaller excitation frequencies. Furthermore, as the gap
shown in Fig. 8; the frequency-amplitude plot of the uncon- width is decreased, the resonance frequency band increases and
strained cantilever is plotted as well for better comparison. As the peak amplitude occurs at larger excitation frequencies.
seen, due to increased stiffness coefficient, the change in the slope Figure 10 shows the effect of the base-excitation amplitude on
of the frequency-response becomes sharper when the cantilever the nonlinear frequency-amplitude response of the constrained canti-
hits the motion constraint. Additionally, the frequency band of lever. It is seen that at a small base-excitation amplitude of 0.005, the
resonance increases with increasing stiffness coefficient. It is cantilever does not hit the motion constraint, hence displaying a
interesting to note that the resonance response of the constrained smooth response. At larger base-excitation amplitudes, on the other
cantilever changes dramatically when K1 is increased from hand, the cantilever does hit the motion constraint causing a sharp
2.0  102 to 2.0  103; however, a smaller change in the resonance change in the slope of the resonance response. As seen, by increasing
response is observed as K1 is increased from 2.0  103 to the base-excitation amplitude, the resonance frequency band increases
2.0  104. Increasing the stiffness coefficient even further, to and the peak amplitude occurs at larger excitation frequencies.
6.0  104, causes only very slight change in the resonance The frequency-amplitude plots of the constrained cantilever for
response, indicating convergence. Hence, K1 ¼ 2.0  104 can be two locations of the motion constraint are illustrated in Fig. 11.
selected as the stiffness of the constraint since it gives a reasona- As seen, as the motion constraint is moved slightly to the right
ble approximation of the motion constraint while allowing for (i.e., further away from the base), the cantilever hits it at smaller
smooth numerical simulations. excitation frequencies. Furthermore, the peak amplitude occurs at
The effect of the gap width (clearance) of the motion constraint larger excitation frequencies as the constraint is moved to the right
on the nonlinear resonance response of the constrained cantilever on the x-axis.
is depicted in Fig. 9. As seen, as the gap width is decreased, the The effect of the material damping coefficient gd on the reso-
cantilever hits the motion constraint at smaller oscillation ampli- nance response of the constrained cantilever is shown in Fig. 12.
tudes; as a result, the change in slope of the resonance response As seen, as gd is increased, the multivalued resonance region

Journal of Applied Mechanics MARCH 2019, Vol. 86 / 031001-9


Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020
Fig. 12 Effect of material damping coefficient on the frequency-amplitude plots of the constrained canti-
lever: (a) maximum tip transverse displacement and (b) maximum tip longitudinal displacement. The
solid line showing stable solution, while the dotted line showing unstable one. g0 5 0.03 and
K1 5 2.0 3 104, and w0 5 0.018.

becomes wider and the peak amplitude occurs at smaller excita- results in widened resonance frequency band. Furthermore, it is
tion frequencies. It is interesting to note that increasing gd from shown that as the clearance between the motion constraint and the
0.006 to 0.007 causes a further reduction in the amplitude com- cantilever is increased, the change in slope of the frequency-
pared to the case when gd is increased from 0.005 to 0.006. response, due to hitting the motion constraint, occurs at smaller exci-
tation frequencies. Examining the effect of the base-excitation ampli-
4 Conclusions tude on the nonlinear resonance response of the system shows that at
small base-excitation amplitudes the cantilever does not hit the
The nonlinear large-amplitude oscillation of a cantilever under motion constraint; as the base-excitation amplitude is increased to
base excitation has been examined in the presence of motion- larger amplitudes, the cantilever hits the motion constraint causing a
limiting constraints on both sides of the cantilever. Hamilton’s change in the slope of the frequency-amplitude. Additionally, due to
principle is employed to derive the equation governing the rota- increased base-excitation amplitude, the resonance frequency band
tional motion of the cantilever centerline, based on the increases and the peak amplitude occurs at larger excitation frequen-
Euler–Bernoulli beam theory and centerline inextensibility assump- cies. If the motion constraint is moved further away from the
tion. The equation of motion is discretized through use of the Galer- clamped end, the cantilever hits it at smaller excitation frequencies
kin technique; the trigonometric functions in the equation of and the resonance peak occurs at larger excitation frequencies.
motion are kept intact to ensure accurate results. The discretized
equations are solved making use of a continuation technique, while
examining the effect of various system parameters.
Appendix: Verification and Convergence
It is shown that unlike other nonlinear models of the cantilever In order to verify the model and numerical results of the present
(based on transverse motion), the present model can predict large- study, a nonlinear static analysis is conducted on a cantilever with-
amplitude oscillations even when the tip angle becomes larger out a stopper and the obtained results are compared to three-
than p/2. It is shown that the presence of a motion constraint dimensional (3D) finite element analysis (FEA) results obtained
reduces the amplitude of the oscillation and widens the band of using ABAQUS/CAE. In order to examine very large-amplitude dis-
resonance frequency vibration. Examining the effect of the stiff- placements, the cantilever is assumed to be under a concentrate
ness of the motion constraint revealed that, as the stiffness is tip load, always perpendicular to the beam. For such a system,
increased, the change in the slope of the frequency-response the nonlinear dimensionless static equation of motion can be
becomes sharper when the cantilever hits the motion constraint; this derived as

Fig. 13 Comparison between the cantilever tip deflections under perpendicular tip load: (a) transverse
deflection and (b) longitudinal deflection. Solid line: present study; symbols: 3D FEA.

031001-10 / Vol. 86, MARCH 2019 Transactions of the ASME


Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020
Fig. 16 Effect of number of degrees-of-freedom on the
frequency-amplitude diagrams of the system
Fig. 14 Deformed configuration of the cantilever at various
forcing amplitudes. Solid line: present study; symbols: 3D FEA.

configuration of the cantilever under various tip load magnitudes


is plotted in Fig. 14 for the two models, again showing very simi-
lar predictions by the two models; it should be noted that x and z
are dimensionless with respect to the length of the cantilever.
Hence, the comparisons in Figs. 13 and 14 verify the developed
model and numerical technique in the present study.
Furthermore, Fig. 15 shows the maximum strain in the cantile-
ver as a function of x when fs ¼ 7.68. In particular, the strain is
calculated at z ¼ h/2, where it is maximum for the system under
consideration, and plotted as a function of x. Referring to Fig. 14,
it is seen that for fs ¼ 7.68, the system undergoes very large-
amplitude deformation. However, as seen in Fig. 15, even for
such large deformation, the strains remain small.
A convergence analysis is shown in Fig. 16 showing a compari-
son between the frequency-amplitude diagrams of the system
obtained using 5DOF and 6DOF discretized models of the system.
As seen, both models predict almost the same response showing
that taking into account five modes ensures converged results.

References
[1] Liu, H., Lee, C., Kobayashi, T., Tay, C. J., and Quan, C., 2012, “Piezoelectric
Fig. 15 Strain distribution as a function of x at z 5 2h/2 for MEMS-Based Wideband Energy Harvesting Systems Using a Frequency-
cantilever under static load fs 5 7.68 (the case with the largest Up-Conversion Cantilever Stopper,” Sens. Actuators A: Phys., 186, pp.
242–248.
force in Fig. 14)
[2] Huicong, L., Chengkuo, L., Takeshi, K., Cho Jui, T., and Chenggen, Q., 2012,
“Investigation of a MEMS Piezoelectric Energy Harvester System With a
Frequency-Widened-Bandwidth Mechanism Introduced by Mechanical
" ð1 ð1 # Stoppers,” Smart Mater. Struct., 21(3), p. 035005.
@2h [3] Moon, F. C., and Shaw, S. W., 1983, “Chaotic Vibrations of a Beam With Non-
þ fs cos h ðdD ð x  1Þcos hÞdx þ sin h ðdD ð x  1Þsin hÞdx Linear Boundary Conditions,” Int. J. Non-Linear Mech., 18(6), pp. 465–477.
@x2 x x [4] Shaw, S. W., and Holmes, P. J., 1983, “A Periodically Forced Piecewise Linear
¼0 (15) Oscillator,” J. Sound Vib., 90(1), pp. 129–155.
[5] Choi, Y. S., and Noah, S. T., 1988, “Forced Periodic Vibration of Unsymmetric
in which fs ¼ fL2 =ðEIÞ with f being the forcing amplitude Piecewise-Linear Systems,” J. Sound Vib., 121(1), pp. 117–126.
[6] Heiman, M. S., Bajaj, A. K., and Sherman, P. J., 1988, “Periodic Motions and
(dimensional). Bifurcations in Dynamics of an Inclined Impact Pair,” J. Sound Vib., 124(1),
Using the same discretization procedure and numerical tech- pp. 55–78.
nique introduced in Sec. 2, the nonlinear static response of the [7] Shaw, S. W., and Rand, R. H., 1989, “The Transition to Chaos in a Simple
cantilever is obtained numerically. The cantilever nonlinear static Mechanical System,” Int. J. Non-Linear Mech., 24(1), pp. 41–56.
[8] Li, G. X., Rand, R. H., and Moon, F. C., 1990, “Bifurcations and Chaos in a
deflection is also obtained using ABAQUS/CAE, using continuum Forced Zero-Stiffness Impact Oscillator,” Int. J. Non-Linear Mech., 25(4), pp.
shell elements. The results obtained by the present study are com- 417–432.
pared to those obtained through 3D FEA and plotted in Figs. 13 [9] Natsiavas, S., 1990, “On the Dynamics of Oscillators With Bi-Linear Damping
and 14. The transverse and longitudinal displacements of the can- and Stiffness,” Int. J. Non-Linear Mech., 25(5), pp. 535–554.
[10] Natsiavas, S., 1993, “Dynamics of Multiple-Degree-of-Freedom Oscillators
tilever tip are shown in Fig. 13 for the two models. As seen, the With Colliding Components,” J. Sound Vib., 165(3), pp. 439–453.
model developed in this study predicts almost the same displace- [11] Nigm, M. M., and Shabana, A. A., 1983, “Effect of an Impact Damper on a
ments as the nonlinear 3D FEA. Additionally, the deformed Multi-Degree of Freedom System,” J. Sound Vib., 89(4), pp. 541–557.

Journal of Applied Mechanics MARCH 2019, Vol. 86 / 031001-11


[12] Bj€orkenstam, U., 1977, “Impact Vibration of a Bar,” Int. J. Mech. Sci., 19(8), Harvester for Low-Frequency and Wide-Bandwidth Operation,” Sens. Actua-
pp. 471–481. tors A: Phys., 208, pp. 56–65.
[13] Masri, S., Mariamy, Y., and Anderson, J., 1981, “Dynamic Response of a Beam [18] Wu, Y., Badel, A., Formosa, F., Liu, W., and Agbossou, A., 2014, “Nonlinear
With a Geometric Nonlinearity,” ASME J. Appl. Mech., 48(2), pp. 404–410. Vibration Energy Harvesting Device Integrating Mechanical Stoppers Used as
[14] Metallidis, P., and Natsiavas, S., 2000, “Vibration of a Continuous System Synchronous Mechanical Switches,” J. Intell. Mater. Syst. Struct., 25(14), pp.
With Clearance and Motion Constraints,” Int. J. Non-Linear Mech., 35(4), pp. 1658–1663.
675–690. [19] Nayfeh, A. H., and Pai, P. F., 2008, Linear and Nonlinear Structural Mechan-
[15] Liu, S., Cheng, Q., Zhao, D., and Feng, L., 2016, “Theoretical Modeling and ics, Wiley, Hoboken, NJ.
Analysis of Two-Degree-of-Freedom Piezoelectric Energy Harvester With [20] Farokhi, H., Ghayesh, M. H., and Hussain, S., 2016, “Large-Amplitude Dynam-
Stopper,” Sens. Actuators A: Phys., 245, pp. 97–105. ical Behaviour of Microcantilevers,” Int. J. Eng. Sci., 106, pp. 29–41.
[16] Soliman, M. S. M., Abdel-Rahman, E. M., El-Saadany, E. F., and Mansour, R. [21] Abdalla, M. M., Reddy, C. K., Faris, W. F., and G€ urdal, Z., 2005, “Optimal
R., 2008, “A Wideband Vibration-Based Energy Harvester,” J. Micromech. Design of an Electrostatically Actuated Microbeam for Maximum Pull-In
Microeng., 18(11), p. 115021. Voltage,” Comput. Struct., 83(15–16), pp. 1320–1329.
[17] Halim, M. A., and Park, J. Y., 2014, “Theoretical Modeling and Analysis of [22] Allgower, E. L., and Georg, K., 2003, Introduction to Numerical Continuation

Downloaded from https://asmedigitalcollection.asme.org/appliedmechanics/article-pdf/86/3/031001/6087702/jam_086_03_031001.pdf by University Of Toronto Library user on 29 July 2020


Mechanical Impact Driven and Frequency Up-Converted Piezoelectric Energy Methods, Society for Industrial and Applied Mathematics, Philadelphia, PA.

031001-12 / Vol. 86, MARCH 2019 Transactions of the ASME

You might also like