Anda di halaman 1dari 23

Downloaded from rspa.royalsocietypublishing.

org on October 4, 2010

Designs for an adaptive tuned vibration absorber


with variable shape stiffness element
Philip Bonello, Michael J Brennan, Stephen J Elliott, Julian F.V Vincent and George
Jeronimidis
Proc. R. Soc. A 2005 461, 3955-3976
doi: 10.1098/rspa.2005.1547

References This article cites 11 articles


http://rspa.royalsocietypublishing.org/content/461/2064/3955.full.
html#ref-list-1

Email alerting service Receive free email alerts when new articles cite this article - sign up in
the box at the top right-hand corner of the article or click here

To subscribe to Proc. R. Soc. A go to: http://rspa.royalsocietypublishing.org/subscriptions

This journal is © 2005 The Royal Society


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Proc. R. Soc. A (2005) 461, 3955–3976


doi:10.1098/rspa.2005.1547
Published online 4 October 2005

Designs for an adaptive tuned vibration


absorber with variable shape stiffness element
B Y P HILIP B ONELLO 1 †, M ICHAEL J. B RENNAN 1 , S TEPHEN J. E LLIOTT 1 ,
J ULIAN F. V. V INCENT 2 AND G EORGE J ERONIMIDIS 3
1
Institute of Sound and Vibration Research, University of Southampton,
Southampton SO17 1BJ, UK
(philip.bonello@manchester.ac.uk)
2
Centre for Biomimetics and Natural Technologies, University of Bath,
Bath BA2 7AY, UK
3
Centre for Biomimetics, University of Reading, Reading RG6 6AH, UK

An adaptive tuned vibration absorber (ATVA) with a smart variable stiffness element is
capable of retuning itself in response to a time-varying excitation frequency, enabling
effective vibration control over a range of frequencies. This paper discusses novel methods
of achieving variable stiffness in an ATVA by changing shape, as inspired by biological
paradigms. It is shown that considerable variation in the tuned frequency can be achieved
by actuating a shape change, provided that this is within the limits of the actuator.
A feasible design for such an ATVA is one in which the device offers low resistance to the
required shape change actuation while not being restricted to low values of the effective
stiffness of the vibration absorber. Three such original designs are identified: (i) A pinned–
pinned arch beam with fixed profile of slight curvature and variable preload through an
adjustable natural curvature; (ii) a vibration absorber with a stiffness element formed from
parallel curved beams of adjustable curvature vibrating longitudinally; (iii) a vibration
absorber with a variable geometry linkage as stiffness element. The experimental results
from demonstrators based on two of these designs show good correlation with the theory.
Keywords: vibration absorber; vibration control; smart structures; biomimetics

1. Introduction

The tuned vibration absorber (TVA) has been used for many years since its
inception by Ormondroyd and Den Hartog (1928). The device is an auxiliary
spring–mass–damper system whose natural frequency (‘tuned frequency’) is tuned
to suppress the vibration at its point of attachment to a host structure. The TVA
can be used in two distinct ways, depending on the application (von Flotow et al.
1994). When used as a ‘mass damper’, the TVA is tuned so as to dampen the
contribution of a particular mode of the host structure to the vibration over a wide
range of excitation frequencies. In this application, an optimal damping level of

† Present address: School of Mechanical, Aerospace, and Civil Engineering, University of


Manchester, PO Box 88, Sackville Street, Manchester M60 1QD, UK.

Received 11 March 2005


Accepted 4 July 2005 3955 q 2005 The Royal Society
Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3956 P. Bonello and others

(a) (b)
me me

complete K(1+jh) effective K(1+jh)


TVA TVA

w = wt= K me
host
structure

f (t) = Re(Fe jw t)

Figure 1. Tuned vibration absorber used as a vibration neutraliser.

damping is required in the stiffness element. When used as a ‘vibration neutraliser’,


the TVA is tuned so as to cancel the vibration at a particular forcing frequency.
Total suppression of the vibration at this frequency is achieved when there is no
damping in the stiffness element. Deviation from the tuned condition (mistuning)
degrades the performance of either variant of the TVA (von Flotow et al. 1994) and
it was shown in Brennan (1997) that a mistuned vibration neutraliser could
actually increase the vibration of its host structure. To avoid mistuning, smart or
adaptive tuned vibration absorbers (ATVAs) have been developed. Such devices
are capable of retuning themselves in real time and an overview of some of these is
given in Brennan (2000) and von Flotow et al. (1994). Adaptive technology is
especially important in the case of the vibration neutraliser since the low damping
requirement in the stiffness element can raise the host structure vibration to
dangerous levels in the mistuned condition.
This paper focuses exclusively on the ATVA used as a vibration neutraliser.
For excitation frequencies below its first natural frequency as a free body, any
auxiliary system used as a vibration absorber can be reduced to an equivalent
two degree of freedom system as shown in figure 1a, where me, K and h are
the effective mass, stiffness and loss factor of the device, respectively. When
the device is attached to its host structure, only the upperffi mass is effective
pffiffiffiffiffiffiffiffiffiffiffiffi
(figure 1b). The tuned frequency of the device is ut Z K=me rad sK1. For zero
ATVA damping, optimal tuning is achieved when utZu where u is the
excitation frequency; when damping is taken into account, this tuning condition
is near optimal. The resulting vibration attenuation is defined as the ratio of the
vibration amplitude at the point of attachment without the ATVA to the
vibration amplitude with the ATVA attached.
At the heart of an ATVA is a stiffness element whose stiffness can be adjusted
in real time so as to maintain the tuned condition in the situation when u and/or
ut may drift. To maximize vibration attenuation, the stiffness element should
have low structural damping (represented by h in figure 1b) and any mechanism
used for stiffness adjustment should add as little as possible to the redundant
mass of the device (Brennan 1997; Kidner & Brennan 1999). Of course, the
stiffness element should also be tunable over an adequate range of frequencies,
and the adjustment should be rapid and with minimum power requirement and
the device must be cheap and easy to manufacture. The technical challenge is to
design a stiffness element with such attributes.

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3957

Various designs for variable stiffness elements in ATVAs have been proposed
(e.g. Hong & Ryu 1985; Longbottom et al. 1990; Franchek et al. 1995; Carneal et al.
2004). Rustighi et al. (2005) recently achieved variable stiffness by using a stiffness
element made of a shape memory alloy, whose Young modulus changes
with temperature, and so can be controlled by adjusting the electrical current
through it. Such a device is easy to manufacture, requiring no actuating mechanism
for stiffness adjustment. However, the maximum variability of the tuned frequency
of such a device is typically found to be around 20% and, for the device in Rustighi
et al. (2005), it took 2 min with a 9 A current to achieve this change. Hence, such a
device is only suitable for applications where any potential mistuning is gradual.
Another strategy for stiffness adjustment involves controlling the shape of the
stiffness element. One such device, first proposed by Walsh & Lamancusa (1992),
had a stiffness element composed of a pair of beams with adjustable gap between
them. This design was modified by Brennan (2000) and developed by Kidner and
Brennan (2002), who used a servomotor to adjust the gap automatically.
A maximum adjustment of 35% in tuned frequency was achieved and the
response was much more rapid than that in Rustighi et al. (2005). However, the
actuating mechanism was still relatively slow to operate and added considerably
to the redundant mass of the device, thereby degrading the vibration
attenuation.
This paper explores the concept of shape control for achieving variable
stiffness in an ATVA, presenting novel ideas inspired by biological paradigms:
changing curvature to adjust stiffness, as observed in plant stems and leaves
(Vincent & Jeronimidis 1991), and controlling stiffness by adjusting geometry, as
in musculo-skeletal structures like the flight motor of an insect (Brennan et al.
2003). Variable curvature is considered first, followed by variable geometry.
Effective designs based on these principles are then identified and results from
two experimental demonstrators, based on variable curvature and variable
geometry, respectively, are presented.

2. Variation of stiffness due to curvature change

The aim of this section is to consider some designs for ATVAs using the curvature
control concept and to discuss how this could be realized.

(a ) ATVA with a shallow pinned arch in lateral vibration


Consider the shallow arch of homogeneous material with boundary conditions
as shown in figure 2a where KS is the stiffness of the longitudinal end restraints.
The system is shown unstressed in figure 2a and in figure 2b the beam is
preloaded with an axial load P0 (P0O0 for compression). Since the beam is
pivoted at its ends, it is assumed that the entire deformation is due to bending.
The stressed and unstressed profiles are, respectively, given by
wi Z H sin ðpx=LÞ; ð2:1aÞ

wP0 Z HP0 sinðpx=LÞ; ð2:1bÞ


where (Ryder 1986)
HP0 Z H =ð1KP0 =PEL Þ; ð2:2Þ

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3958 P. Bonello and others

Figure 2. Dynamics of pinned shallow arch with end longitudinal restraints of finite stiffness.

and PEL Z p2 EI =L2 is the Euler critical load. Equations (2.1) and (2.2) and the
subsequent analysis assume the arch is shallow, i.e. ðwi0 Þ2 ; ðwP0 0 Þ2 / 1; where
ð Þ 0 Z dð Þ=dx. For small vibrations vZ wKwP0 about the preloaded configuration
wP0 ðxÞ (figure 2c), the neutral surface of the beam is strained by an amount
depending on the stiffness KS of the longitudinal end restraints. In reference
Bonello et al. (2004) it is shown that, for vibrations symmetrical about the mid-
section, the equation of motion is given by
ÐL 0 0
iv 00 00 0 wP0 v dx
EIv C v P0 KwP0 C gv€ Z 0; ð2:3Þ
2f1=2KB Þ C 1=KS g
where E, I, A, g are the Young modulus, second moment of area of cross-section
about neutral axis, cross-sectional area and mass per unit length, respectively, of
the curved beam and KBZEA/L, the axial stiffness of the straightened beam.
Equation (2.3) neglects the effects of inertia in the longitudinal direction. This
equation is a generalization of the case considered in (Reissner 1954) of a pinned–
pinned shallow arch with full end longitudinal restraint and no preload. In that
work, it was shown that the longitudinal inertia was negligible. Equation (2.3)
also applies for vibration that is not symmetric about the mid-section, provided
that KSZN. If vibration in the first symmetric mode is considered, the response
is given approximately by
v Z A sin ðpx=LÞ sin u1 t: ð2:4Þ
Substituting equations (2.4) and (2.1b) in equation (2.3) and using equation (2.2)
gives the equation governing the tuned frequency of a potential TVA consisting

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3959

350

% increase in natural frequency


analytical, equation (2.6)
300
finite element
250
200
150
100
50

0 1 2 3 4
H h

Figure 3. Variation of natural frequency with curvature of a shallow pinned arch of rectangular
section 2h with no preload and KSZN.

of a preloaded pinned shallow arch:


   
H 1 HP0
u2t Z u2s C =f1 C 2KB =KS g ; ð2:5Þ
HP0 2 r
where us rad sK1 is the fundamental frequency of lateral vibrations of the
straightened beam and r is the radius of gyration of cross-section about the
neutral axis. All examples in pffiffiffithis section refer to a rectangular section of
thickness 2h for which r Z h= 3.
Equation (2.5) shows that, for P0Z0 (0H Z HP0 ) and KSZ0 (no preload, no
end longitudinal restraint), the dependency of u2t on curvature disappears.
If P0Z0, KSZN (no preload, full end longitudinal restraint), equation (2.5)
reduces to the expression derived by Reissner (1954) for the first symmetrical
mode of a pinned–pinned shallow arch with full end restraint and no preload:
  
2 2 1 H 2
ut jP0Z0;KSZN Z us 1 C : ð2:6Þ
2 r
Figure 3 shows the percentage increase in u2t jP0Z0;KSZN with curvature under such
conditions for an arch for which 2hZ0.75!10K3 m, the width (normal to plane
of paper) is 15 mm, and LZ0.15 m. The theoretical result given by equation (2.6)
compares very well with the result obtained by finite element (FE) analysis of the
arch using shell elements up to at least H/hZ3.5. While the variation of
frequency with curvature in equation (2.6) has been shown to be considerable, it
has little practical use for an ATVA since it applies only for P0Z0 and it would
be extremely difficult to adjust the curvature H of the pinned–pinned beam in
real time without loading it axially. The effect of the preload will henceforth be
considered.
Equation (2.5) gives two methods for varying ut in the ATVA: (i) for fixed H,
vary HP0 ; (ii) for fixed HP0 , vary H. In either of these cases, it is preferable to
have full longitudinal end restraint in vibration (KSZN) so that the relevant
equation to consider is   
2 2 H 1 HP0 2
ut Z us C : ð2:7Þ
HP0 2 r

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3960 P. Bonello and others

It is assumed that the vibration of the TVA is symmetrical about the mid-section
of the arch. However, in practice, due to inherent slight asymmetry in conditions,
an antisymmetric mode of vibration will also be excited, becoming significant
when the frequency given by equation (2.7) becomes close to the frequency
u2 rad sK1 of the first antisymmetric mode of the pinned–pinned beam. In this
case the TVA will not reduce to the simple model of figure 1b and will cease to
function effectively. Substituting vZ A sin ð2px=LÞ sin u2 t in equation (2.3) with
KSZN, along with the expression for the deformed profile about which vibration
occurs (equation (2.1b)), gives, after some algebra:
u22 Z u2s ð16K4P0 =PEL Þ Z u2s ð12 C 4H =HP0 Þ: ð2:8Þ
(by substituting for P0/PEL from equation (2.2)). The condition u2t ! u22 limits
the variation of H and HP0 in the ATVA since, from equations (2.7) and (2.8):
ðHP0 =rÞ3 K24ðHP0 =rÞK6ðH =rÞ! 0 for u2t ! u22 : ð2:9Þ
The methods (i) and (ii) for varying ut are now considered in turn.

(i) Varying ut by changing deformed shape


For a given natural (undeformed) curvature (fixed H ), HP0 is varied by
moving end supports inwards or outwards (figure 4a). In this case, a finite
stiffness KS in equation (2.5) would reduce the variability in ut. Hence, full
longitudinal restraint in vibration is required. An example of an actuator that
fulfils this requirement is a motor driven screw. Figure 5a shows the variation
obtained in ut with variation in HP0 =h for three different values of the
undeformed condition H/h. Figure 5b shows the corresponding force requirement
of the actuator. The upper limit on HP0 =h for a range of H/h (see equation (2.9))
is given by the solid line in figure 5c. The dashed line in figure 5c definespthe ffiffiffiffiffi
undeformed condition and intersects with the solid line at HP0 =hZ H =hZ 10,
which is the limiting condition for the no-preload (Reissner 1954). Figure 5 shows
that, while considerable variability in ut is possible, considerable amount of force
is required of the actuator due to the high longitudinal stiffness of the curved
beam. Moreover, the actuator will be a redundant mass that degrades the
vibration attenuation. It is also evident from figure 5c that there are problems of
antisymmetric mode interference.

(ii) Varying ut by changing natural curvature


For a given curvature about which vibration occurs (fixed HP0 ), the
undeformed curvature is varied (i.e. vary H ). This can be achieved in practice
by bonding a layer of piezo–ceramic (figure 4b) and applying different levels of
voltage to the piezo. If there is full longitudinal end restraint (KSZN), HP0 will
remain approximately the same but the preload P0 will be altered, thereby
adjusting the tuned frequency. It is understood that a piezo-actuated beam is
composite, however, it is reasonably assumed that the above equations,
developed for a homogeneous section, give a fair description of the behaviour
of the practical application. Figure 6a shows the variation obtained in ut with
variation in H/h for three different values of the deformed condition HP0 =h, and
figure 6b shows the corresponding variation in preload. The lower limit on H/h

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3961

(a) (b)
slight adjustment in distance fixed distance

HP HP
0 0

actuator rigid cross member piezo bonded to beam


attach to structure attach to structure
fixed H, variable HP fixed HP , variable H
0 0

Figure 4. Two alternative designs for a shallow arch ATVA.

(a) 200 (b) 1


H/h=1
% increase in tuned frequency

0
150 H/h=2 –1
H / h = 2.5 –2
100
–3
P0 H/h=1
–4
50
PE L –5 H/h=2
0 –6
H / h = 2.5
–7
– 50 –8
–9
–100 – 10
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
HP h HP h
0 0

(c) 5

HP 3
0

h
2

1 upper limit on HP /h (equation (2.9))


0
H = HP
0

0
0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
H h

Figure 5. Performance curves for shallow arch ATVA with fixed H, variable HP0 (figure 4a).

over a range of HP0 =h (equation (2.9)) is given by the solid line in figure 6c, where
the dashed line represents
pffiffiffiffiffithe undeformed condition and intersects with the solid
one at HP0 =hZ H =hZ 10 as before. It is seen from figure 6a that this technique
works very well for slight curvatures HP0 =h. Moreover, for such slight curvatures,
there are no limits on the variation of H/h, as far as antisymmetric mode
interference is concerned, since the lower limit on H/h is negative in figure 6c.
Unlike the technique in the previous method §2a(i), the preload P0 is provided by
the support reaction, i.e. the actuator does not act against the high longitudinal
stiffness of the beam. Moreover, the actuation would be incorporated into the
beam making the design simple, cheap and efficient.

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3962 P. Bonello and others

(a) 120 (b) 1


% increase in tuned frequency HP /h= 1 0
100 0
HP /h= 2 –1
80 0
–2
HP /h=2.5
60 0 –3
P0 –4
40 PEL –5
20 HP /h= 1
–6 0
0 –7 HP /h= 2
0
–8 HP /h=2.5
– 20
–9 0
– 40 – 10
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
H h H h

(c) 50
40 lower limit on H / h (equation (2.9))
HP = H
30 0

H
h 20

10

–10
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
HP h
0

Figure 6. Performance curves for shallow arch ATVA with fixed HP0 , variable H (figure 4b).

It is worth pointing out that either device depicted in figure 4 is limited by its
small effective mass. However, this limitation can be overcome to some extent by
the addition of a dense mass at the apex.

(b ) Anticlastic curvature control


Another design considered was a cantilever beam vibration neutraliser whose
tuned frequency was to be controlled by varying the curvature of its cross-
section, i.e. by introducing an anticlastic-type curvature, thereby varying its
flexural rigidity (see figure 7). This is indeed how many leaves vary their effective
bending stiffness. With reference to figure 7b, which shows the cross-section of a
composite cantilever beam, assuming that the cross-section geometry does not
vary significantly due to the vibration, the fractional increase in tuned frequency
from the flat cross-section condition (subtended angle aZ0) is estimated as
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ft Kft jaZ0 I
Z K1; ð2:10Þ
ft jaZ0 I jaZ0
where I is the equivalent second moment of area of the composite section about
the neutral axis NN (figure 7b):
I Z ðINN Þ1 C C ðINN Þ2 : ð2:11Þ
(INN)1, (INN)2 are second moments of area of the individual layer cross-sections
about the neutral axis NN and C Z E2 =E1 , the ratio of the Young Moduli of the

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3963

(a) (b)

2L
E1
A a/2
R h
neutral axis h

N N

host structure mid-surface


E2 length S

ATVA beam cross-section

Figure 7. Cantilever-type ATVA with anticlastic curvature control.

two layer materials. Upon locating axis NN using elementary composite beam
theory, analytical expressions for (INN)1, (INN)2 are determined from the
geometry of figure 7b (Bonello et al. 2004):
 
ðINN Þ1 Z ðSh3 =8Þ U 3 F1 C UF2 C U K1 F3 C U K1 ð1KDÞ2 F4 ; ð2:12aÞ
3
 3 K1 K1 2

ðINN Þ2 Z ðSh =8Þ V F1 C VF2 C V F3 C V ð1 C DÞ F4 ; ð2:12bÞ
C Z E2 =E1 ; ð2:13aÞ
D Z ð1KC Þ=ð1 C C Þ; ð2:13bÞ
2
F1 Z ð1=2Þðh=SÞ½a C sin aKð8=aÞsin ða=2Þ; ð2:14aÞ
2
F2 Z ð1=2Þðh=SÞ½a C sin aKð16=3Þð1=a sin ða=2Þ; ð2:14bÞ
F3 ZKð4=9Þðh=SÞð1=aÞsin2 ða=2Þ; ð2:14cÞ

F4 Z ð128=9Þðh=SÞ
  4      K2
9 S 6 S 2 2 S
!ð1=aÞ 4 K 2 C1 KD sin2 ða=2Þ; ð2:14dÞ
a h a h a h

U Z ð2=aÞðS=hÞK1; ð2:14eÞ

V Z ð2=aÞðS=hÞ C 1: ð2:14f Þ
For a flat cross-section, the expression in (2.11) reduces to
 

I aZ0 Z S aZ0 h 3 =24 2ð1 C C Þ C 6ð1KDÞ2 C 6Cð1 C DÞ2 : ð2:15Þ
Three methods of changing the cross-section curvature that were considered are
as follows.

(a) Subjecting a bimetallic strip cross-section to a temperature change. While


considerable changes of curvature can be achieved in this way, the response
time would be too long for many applications.

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3964 P. Bonello and others

(a) x (b)
My Ny
y z
Mx
Mx
Nx
Nx

My Ny
laminate exploded view of laminate

Figure 8. Two-layer cross-ply laminate.

(b) Using the stretch-bend effect of cross-ply laminates (Jones 1975). With
reference to figure 8, if the cross-section is a cross-ply laminate consisting of
two orthotropic layers of the same composite material that are oriented such
that their fibres are at 908 to each other, then a tensile force Nx per unit
length along an edge will produce curvature Kx in a plane normal to the edge
according to the equations
" # " #  " #" #
Nx A11 A12 30x B11 0 Kx
Z C ; ð2:16aÞ
Ny A12 A22 30y 0 B22 Ky
" # " #  " #" #
Mx B11 0 30x D11 D12 Kx
Z C ; ð2:16bÞ
My 0 B22 30y D12 D22 Ky

where the axes x and y are aligned with the axes of symmetry of material
properties, 30x, 30y are the longitudinal strains in the mid-surface and Kx, Ky
the curvatures and the matrix terms are defined in Jones (1975). It is shown
in Jones (1975) that the coupling coefficients between stretching and
bending are maximum for given material properties when the two layers are
of equal thickness h.
(c) Using piezo-actuation. This is discussed later, in §2d (ii).

Consider, as an example, approach (b) applied to a composite section


cantilever (figure 7) with section 2 mm!50 mm, length 2LZ0.24 mm, density
rZ1600 kg mK3 and orthotropic material properties E1, E2Z130, 10 GPa
(Young Moduli), y12Z0.3 (Poisson ratio (Jones 1975)). Figure 9a shows the
percentage increase in tuned frequency from the flat section condition, computed
using equations (2.10)–(2.15). Figure 9b shows the tension force required to
achieve the necessary curvature, computed by equations (2.16a) and (2.16b) and
verified at two points by FE. The main assumption in figure 9a is that the cross-
section geometry is taken to be invariant during vibration. This assumption was
tested using FE analysis with shell elements, in which the beam mid-section was
constrained only at its point of attachment A to the structure (see figure 7a). The
assumption of invariant cross-section resulted in a small overestimation of the
tuned frequency (typically around 5%). figure 9a shows that significant changes

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3965

(a) 200 (b) 2.0


analytic (equation (2.16a,b))

force (= 2Nx L) (×10 5 N)


percentage increase in
finite element
150 1.5
tuned frequency

100 1.0

50 0.5

0 0 5 10 15 20 25 30 35 40 45 50
5 10 15 20 25 30 35 40 45
a (deg.) a (deg.)

Figure 9. (a) Tuning characteristic for cantilever type TVA in figure 7 estimated from equations
(2.10)–(2.15) for a 2-layer cross-ply cross-section with E1, E2Z130, 10 GPa, y12Z0.3, 2hZ2!10K3 m,
2LZ0.24 mm; (b) Edge force required to achieve cross-section curvature in figure 9(a).

in tuned frequency are possible, provided that the actuator can provide the
necessary changes to the subtended angle a. However, the slope of the graph in
figure 9b is seen to be enormous, indicating that stretching the composite cross-
section is not a practical way of varying the cross-section.
The main drawback with the design in figure 7 is that significant variations
in the subtended angle a are required over a short length S. Since the curvature
KZa/S, this means that the required curvature changes are too high for method
(b). It is also shown in §2d(ii) that piezo-actuation cannot yield the curvature
changes required for this application. Indeed, of methods (a)–(c) listed above,
(a) appears to be the only viable option for this application.
Hence, for the above reasons the concept of anticlastic curvature control for
variable stiffness in figure 7 was not pursued any further.

(c ) ATVA with parallel curved beams in longitudinal vibration


Equation (2.5) shows that the natural frequency of lateral vibrations of a
curved beam would be insensitive to curvature if there were no longitudinal
end restraints. However, its static stiffness in the longitudinal direction (i.e.
along the line joining its ends) does alter with curvature, and the addition of
significant longitudinal inertia in the form of a lumped mass at the end would
clearly make the modal parameters of the resulting system sensitive to
curvature. This is the rationale of the ATVA depicted in figure 10a, where the
stiffness element is composed of a pair of parallel identical beams of adjustable
curvature vibrating longitudinally. Referring to figure 10b, which shows the
geometry of the circular profile of the undeformed beam, assuming small
deformations, the non-dimensional static stiffness of each beam along the
dashed line can be obtained by Castigliano’s theorem:
2a3
K~ Z 2 ; ð2:17Þ
p f2aK3 sin a C a cos ag
P=PE
where K~ Z Z P~ u;
~
u=S

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3966 P. Bonello and others

(a) mass (b)

curved beam curved beam H a


S

attach to structure
vibration absorber beam geometry

Figure 10. Vibration absorber with stiffness element formed from parallel curved beams.

where P, u are the compressive force and deflection, respectively, and PE Z


p2 EI =S 2 is the Euler buckling load of a straight beam of the same length S,
the Young modulus E and second moment of area of cross-section I as the
curved beam. Since the crown height H can be related to the subtended angle a, it
can be shown graphically that K~ zð2=p2 ÞðH =SÞK2 . Hence, since the tuned
~ it is approximately
frequency ft of the TVA is proportional to the square root of K,
inversely proportional to the crown height H of the beams. One way of changing
the curvature of the beams of such an ATVA is to bond a layer of piezoelectric
ceramic to each beam and adjusting the level of DC voltage applied to these
actuators.
The above analysis neglects the nonlinearity in the P–u relationship for finite
deformations. This is quantified in (Bonello 2004) and shown not to affect
significantly the ATVA performance. The analysis also neglects the effect of the
inertia of the beams. In fact, the maximum frequency the ATVA can be tuned to
in practice will be the fundamental frequency of lateral vibration of each beam
due to its own elasticity and inertia.

(d ) Methods of changing curvature


In §2b it was observed that composites and piezo-actuators are unsuitable for
the application described in figure 7. However, they are extremely useful for
those designs of ATVA which require a lower degree of curvature variation for a
given change in natural frequency, as in the proposed designs of figure 4b and
figure 10a. This section discusses how these materials can be used to control
curvature.

(i) Composites
One way of implementing the stretch-bend effect in composites is to use
hydraulic pressure to generate the required tension force along the edge.
Figure 11 shows an initially straight clamped-free beam in which the upper and
lower surfaces are orthotropic laminae of the same composite material and are
oriented such that their fibres are at 908 to each other. The two laminae are

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3967

0.01 m
0.12 m

0.01 m

Figure 11. Finite element model of clamped-free beam with a pressurised fluid-filled cavity
separating a pair of laminae of orthotropic material with material axes at right angles to each other
(thickness of each lamina is 0.25 mm).

separated by a fluid filled cavity. Fibre reinforcements joining the upper and
lower surfaces prevent them from bulging. Figure 11 shows the FE model of the
deformation resulting from the application of a 1 bar pressure. For the
dimensions shown, and a Poisson ratio (y12Z0.3, the tip deflections are
0.13 mm, 0.145 mm, 0.274 mm, respectively, for major and minor Young moduli
(E1, E2) of (130 GPa, 10 GPa), (10 GPa, 5 GPa), (100 GPa, 5 GPa). If such a
beam were incorporated into the devices of figure 4 and figure 10 then these
deflection values would be the changes in the crown height H for a 1 bar change
in pressure. Hence, it is expected that a pressure increase of, say, 20 bar would
result in significant variation to the tuned frequency.
The beam in figure 11 is somewhat analogous to the stem of a plant
(Vincent & Jeronimidis 1991). As in the present case, the turgor pressure in a
stem (typically 20 bar; Vincent & Jeronimidis 1991) is uniform. However, the
bending of the stem is due to the variation of cell wall thickness across the cross-
section, resulting in differential cell wall expansion, and the curvature changes
measured in (Vincent & Jeronimidis 1991) were considerably higher than those
obtained with the composite model of figure 11.

(ii) Piezo-actuation
A piezo-actuator bends a beam to which it is bonded by applying a moment.
The piezo-patch(es) can be placed on both parallel surfaces of the beam with
opposing polarities or one surface only. With one-sided piezo-actuation there is a
tensile force on the beam mid-surface as well. Hence, it is worthwhile
investigating whether the use of a composite beam with a piezo-actuator on
one surface would have any advantage with respect to curvature generation.
If 3p is the longitudinal expansion coefficient (‘free strain’) of the actuator
under a voltage V:
3p Z d31 V =ta ; ð2:18Þ
where d31 is the piezo-electric constant (Srinivasan & Michael McFarland 2001).
With reference to figure 12, following an analysis broadly similar to that in
Srinivasan & Michael McFarland (2001), it can be shown that the piezo will
cause the initially straight beam to bend to a radius R (measured from the beam
mid-surface), with the ratio r of non-dimensional curvature to actuator free

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3968 P. Bonello and others

S
ta piezo-actuator
h E1
beam
h E2

beam mid-surface

Figure 12. Beam with piezo-actuator.

strain ratio being given by


h=R
rZ
3p

Z K6Bð1 C 6C C 9C 2 Þj2 K6Bð1 C C Þð1 C 3C Þj2 T
ð2:19Þ
C6B 2 ð1 C C ÞjT 2 C 6B 2 ð1 C 3CÞjT

! ½ð1 C 3C ÞjKBT½ð1 C 14C C C 2 Þj2 C 4Bð1 C C ÞT 2 j
K1
C6Bð1 C 3C ÞTj C 4Bð1 C 7C Þj C B 2 T 2 
where
B Z Ea =E1 ; ð2:20aÞ
C Z E2 =E1 ; ð2:20bÞ

T Z ta =h; ð2:20cÞ

j Z bb h=ðba ta Þ; ð2:20dÞ
and bb, bs are the beam and actuator widths normal to plane of paper. The
formula applies for a composite beam with two layers of equal thickness h. This
was done to simplify the analysis. Moreover, it was thought that the use of equal
layer thickness might maximise the stretch-bend effect induced by the piezo, as
in a cross-ply laminate with two equal thickness layers under tension. figure 13
shows the variation of jrj with T/2 for baZb b and various C for PZT material
with EaZ59.5 GPa and d31Z212!10K12 V mK1 (Brennan et al. 1994). From
figure 13 it appears that there is some advantage in using a composite beam with
the piezo-actuator. The curves in figure 13 show that there is normally optimum
value of T, Topt, which minimizes the electric field requirement for a given
curvature in a beam of given geometry. Hence, there is the possibility of choosing
the materials of a composite beam to reduce this value of Topt. Substituting
rZr(Topt), taZTopth and 1/RZa/S in equations (2.18) and (2.19) and
eliminating 3p gives the minimum voltage required to generate a subtended
angle of a by piezo-actuation in a beam of given thickness 2h and length S :
Topt h 2 a
Vmin Z : ð2:21Þ
rðTopt Þ d31 S
Figure 14 shows Vmin as a function of a for SZ0.05 m and 2hZ0.5!10K3m and
various material properties. It is evident that high voltages are necessary to
actuate significant changes to the subtended angle a over a short length S,

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3969

0.7
0.6
0.5
h/R 0.4
ep 0.3
E 1= E 2 = 10 GPa E 1= E 2 =130 GPa
0.2
E 1 = 10 GPa, E 2 = 5 GPa E 1 =130 GPa, E 2 = 13 GPa
0.1 E 1 = 5 GPa, E 2 = 10 GPa E 1 =13 GPa, E 2 =130 GPa
0
0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
ta /(2h) ta /(2h)

Figure 13. Variation of curvature to free strain ratio with actuator to beam thickness ratio for baZbb
and EaZ59.5 GPa, d31ZK212!10K12 m VK1.

4000

3500

3000

2500
voltage (V)

2000

1500
E 1= E 2 =10 GPa
1000
E 1 =10 GPa, E 2 = 5 GPa
500
E 1 =5 GPa, E 2 = 10 GPa
0 5 10 15 20 25 30 35 40 45 50
a (deg.)

Figure 14. Minimum voltage versus subtended angle a for b aZb b and 2hZ0.5!10K3 m, SZ0.05 m,
EaZ59.5 GPa, d31ZK212!10K12 m VK1.

showing that piezo-actuation is not suitable for anticlastic curvature control


(figure 7). Section 5 illustrates the usefulness of piezo-actuation for actuating the
curvature changes required for the design in figure 10a.

3. Variable stiffness by controlling geometry

Consider the linkage in figure 15a where the rigid links AB, BC are pivoted at
A, B and C and the cantilevers are of flexural stiffness k. This configuration is
based on the simplified model of the flight motor of an insect used in the analysis
of Brennan et al. (2003). The nonlinear expression relating the applied static
force F to y (figure 15b) was derived in Brennan et al. (2003):
(  )
F b y 2 K1=2 y
Z 1K K1 : ð3:22Þ
2kL L L L

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3970 P. Bonello and others

(a) (b) F
B

y0 L L
y
A q0 C q

rigid links

cantilevers
stiffness k

2b
undeformed deformed

Figure 15. Linkage spring.

From this expression, the non-dimensional linear stiffness for small deflections
DyZyKy0 about the stable equilibrium position of figure 15a can be derived:
F=ð2kLÞ 1
K~ z Z K1: ð3:23Þ
Dy=L ðb=LÞ2
Hence, variable stiffness can be achieved by varying the distance between the
cantilevers. Figure 16 shows a schematic diagram of a self-tuning vibration
absorber that uses this concept, its tuned frequency ft being proportional to the
square root of the stiffness given by equation (3.23). Note that in the device
shown in figure 16, the stiffness element is inverted relative to that of the insect
model in Brennan et al. (2003) so that the actuator forms part of the effective
mass of the device, thereby enhancing vibration attenuation (Brennan 1997).
The force required from the actuator to control the stiffness is minimal but the
actuator needs to have a large stroke. A further requirement for the actuator is
that its longitudinal stiffness should be sufficiently high as to prevent variation in
the distance between the cantilevers as a result of vibration-induced deformation.
Hence, a suitable actuating mechanism for the device would be a screw driven by
a servomotor, like that used in Kidner & Brennan (2002). Alternatively, it might
be possible to use the gel actuator developed in associated research (Santulli
et al. 2002) forming part of the same multi-disciplinary project as the research
presented in this paper. This would be in keeping with the musculo-skeletal
analogy behind the device in figure 16, since the gel actuator can be considered as
the engineering analogue of a muscle.
The effectiveness of the device is limited to frequencies for which inertia effects
of the components forming the linkage are negligible. For low values of q0 the
device is prone to ‘snap through’ to its lower stable equilibrium position at high
amplitudes of vibration (Brennan et al. 2003). Moreover, for a given level of
vibration, nonlinear effects are expected to be more prominent at low values of
q0, as shown in figure 17, which compares the nonlinear expression in equation
(3.22) with the linear (tangential) relationship of equation (3.23) for three
different values of q0. It is also seen from figure 17 that, for deflections from the
stable equilibrium position, the spring hardens in tension (FO0) and softens in
compression (F!0).

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3971

adjust distance

mechanism and
absorber mass

cantilevers

b b
q0
rigid link L rigid link
L
structure

Figure 16. Schematics of linkage spring ATVA.

4. Effective ATVA designs

In the previous sections it was shown that, while considerable variation in tuned
frequency can be achieved by actuating a shape change, this will be limited by
the maximum force that can be provided by the actuator. A feasible design for an
ATVA is one in which the device offers low resistance to the required shape
change actuation. Of course, this should not restrict the device to low values of
ATVA effective stiffness (i.e. the stiffness of the equivalent model in figure 1).
Hence, the best designs are those in which the actuator force is uncoupled as
much as possible from the effective stiffness of the TVA. For example, the devices
in figures 4b, 10a and 16 are effective since, although their effective stiffness can
be considerably high, it is controlled with relatively little effort. For the device in
figure 4b, this is achieved by merely altering the natural (undeformed) shape,
which is equivalent to the piezo-actuator acting against the relatively low
flexural stiffness of a free–free beam. In figure 10a the effective stiffness of the
device is controlled by changing the curvature of the beams. While the controlled
stiffness can be quite high, the curvature of each beam is relatively easy to adjust
by flexure since it is pivoted at the ends and one end is free to move. In figure 16
the effective stiffness of the device is controlled by adjusting the distance between
the cantilevers and the (hinged) linkage stiffness element offers no resistance to
this adjustment.

5. Experimental demonstrators

Demonstrators were built for two of the three effective designs identified in the
previous section. This section presents some experimental data obtained from
these devices. The aims are twofold: to prove that they perform as predicted
under different levels of excitation and to illustrate the variability in tuned
frequency obtainable. For each setting of the ATVA, the tuned frequency ft
(in Hz) and the damping loss factor h were measured by applying random
excitation at the base of the device and measuring the transmissibility frequency
response relating the acceleration at the base, a1(t), to that of the absorber mass

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3972 P. Bonello and others

b / L = 0.7
0.15
b / L = 0.8
0.10 b / L = 0.6

0.05
F
0
2kL
– 0.05

-0.10
linearized, equation (3.23)
– 0.15

– 0.5 0 0.5
y/L

Figure 17. Nonlinear relationship between F and y for various b/L values.

a1(t). The measurements were taken for two levels of excitation, the higher one
corresponding to peak values of a2(t) in excess of 10 m sK2. Assuming a linear
relationship between these two signals and a single degree of freedom (SDOF)
model for the effective part of the demonstrator, as in figure 1b:

A2 1 C jh
TZ Z ; ð4:24Þ
A1 1Kðf =ft Þ2 C jh

where h is the damping loss factor of the stiffness element, f is the frequency and
A1 and A2 are the complex
 amplitudes
 of a1(t) and a2(t) for harmonic vibration,
i.e. when ar ðtÞZ Re Ar e j2pft , rZ1, 2. jT j is maximum at fZft and ft and h
were accurately located by considering that, at this point, Re{T}Z1 and
Im{T}ZK1/h.
Figure 18 shows a picture of a demonstrator for the ATVA of figure 10a. Each
curved beam is a piezo-actuator (THUNDER TH 7-R; Face International 1998)
consisting of a piezo–ceramic layer 0.25 mm thick sandwiched between a
0.15 mm thick aluminium layer and a 0.2 mm thick steel plate. The beams are
supported on miniature ball bearings at their ends, allowing freedom of rotation.
The absorber mass is guided in the direction of vibration by a linear bearing
running on the central column that is fixed to the base. The beam curvature was
adjusted by varying the piezo-actuator voltage. The piezo-actuators can take
between K300 and C600 V DC (Face International 1998), however, the
experiments results were restricted to a voltage range of K280 to C480 V DC.
Since there is negligible leakage through the high resistance ceramic, the current
required to maintain a given curvature setting is miniscule.
At low vibration amplitudes, the lowest measured value of ft (at K280 V) was
36 Hz and the highest 56 Hz (at C480 V), with the value of ft at 0 V being 42 Hz.
The corresponding frequencies at seven times the vibration amplitude were only
slightly changed (36, 55 and 41 Hz, respectively). The experimental work on this
demonstrator is described in greater detail in a separate paper (Bonello et al.
2005). Figure 19 shows the percentage variation in ft with applied voltage.

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3973

13 cm
Figure 18. Prototype ATVA with variable curvature beams as stiffness element.

50
percentage increase in tuned frequency

predicted (using equation (2.17))


40 measured (low amplitude)
measured (7 × amplitude)
30

20

10

– 10

– 20
– 30
– 300 – 200 – 100 0 100 200 300 400 500
voltage (V, DC)

Figure 19. Tuning characteristic for prototype ATVA.

The prediction was obtained by using the stiffness–curvature relation of equation


(2.17) and the curvature–voltage relationship derived from manufacturer data,
which was only available for the voltage range K300 to C300 V. As discussed
previously, the inertia of the beams limits the maximum tuned frequency—hence
the divergence between measurement and prediction as the tuned frequency
increases. It is seen from figure 19 that the maximum measured variation in ft is
almost 50% (from K20 to C30%).
Figure 20 shows a laboratory demonstrator for the variable geometry linkage
device of figure 16. Note that the device in figure 20 is an inverted version of that
shown in figure 16. The distance between the cantilevers was manually adjusted
by turning a screw. The nominal configuration was bnom/LZ0.8 (i.e. q0,nomZ378
in figure 16).
For the two extreme settings Db/LZb/LKbnom/LZ0.1 and Db/LZK0.075,
the tuned frequencies are 70 Hz and 150 Hz, respectively. Figure 21a shows the

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3974 P. Bonello and others

Figure 20. Demonstrator for manually tuned linkage spring device.

(a) (b)
10 2
measured
measured
from equation (4.24)
from equation (4.24)
10 1 with measured
with measured
ft and h
ft and h
magnitude

10 0

10 –1

10 – 2

1.0
measured measured
0.5 from equation (4.24) from equation (4.24)
with measured with measured
cos f

0 ft and h ft and h

–0.5

–1.0
0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400
frequency (Hz) frequency (Hz)
low amplitude high amplitude

Figure 21. Transmissibility plots at low amplitude (Db=LZ b=LKbnom =L; bnom =LZ 0:8).

magnitude of the transmissibility and the cosine of its phase f for Db/LZ0.1 at
low amplitude. It is seen that linear SDOF theory holds and cos f z0 when fZft.
Figure 21b show that, at 10 times the amplitude, the linear theory still holds in
the region of the tuned frequency and this frequency changes only slightly. The
strong harmonics that appear, indicating non-linearity, will have no effect on the
operation of the ATVA.

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

Adaptive tuned vibration absorber with variable shape stiffness element 3975

50
predicted

percentge increase in tuned frequency


40 measured (low amplitude)
measured (5 × amplitude)
30 measured (10 × amplitude)
20
10
0
– 10
– 20
– 30
– 40
– 50
– 0.10 – 0.08 – 0.06 – 0.04 – 0.02 0 0.02 0.04 0.06 0.08 0.10
Db / L

Figure 22. Tuning characteristic (Db=LZ b=LKbnom =L; bnom =LZ 0:8).

Figure 22 compares the predicted and measured tuning characteristics for the
nominal configuration bnom/LZ0.8, where the correlation between simulation and
measurement is seen to remain reasonably good at high amplitude. The maximum
variation in ft is 70% (K40 to C30%) for a total actuator movement of 1.6 cm.

6. Conclusions

This paper has explored the concept of shape control for achieving variable
stiffness in an ATVA, presenting novel ideas inspired by biological paradigms
where a change in curvature or geometry is used to control stiffness. Several
approaches were considered: using a shallow arch in transverse vibration, curved
beams vibrating longitudinally, changing anticlastic curvature, and changing the
geometry of a linkage system. A study on the use of composites and piezo-
actuators to achieve the required shape change was also performed. It was
concluded that considerable variation in the tuned frequency can be achieved by
actuating a shape change, provided that this is within the limits of the actuator.
A feasible design for such an ATVA is one in which the device offers low resistance
to the required shape change actuation while not being restricted to low values of
the effective stiffness of the vibration absorber. Hence, the best designs are those in
which the actuator force is uncoupled as much as possible from the effective
stiffness of the absorber. Three such designs were identified: (i) a pinned–pinned
arch beam with fixed profile of slight curvature and variable preload through an
adjustable natural curvature; (ii) a vibration absorber with a stiffness element
formed from curved beams of adjustable curvature in parallel vibrating
longitudinally; (iii) a vibration absorber with a variable geometry linkage as the
stiffness element. The experimental results from demonstrators based on the latter
two designs showed good correlation with the theory.
The authors would like to acknowledge the Engineering and Science Research Funding (EPSRC) of
the UK for funding this work.

Proc. R. Soc. A (2005)


Downloaded from rspa.royalsocietypublishing.org on October 4, 2010

3976 P. Bonello and others

References
Brennan, M. J. 1997 Vibration control using a tunable vibration neutraliser. J. Mech. Eng. Sci.
211, 91–108.
Brennan, M. J. 2000 Actuators for active control—tunable resonant devices. J. Appl. Mech. Eng. 5,
63–74.
Brennan, M. J., Day, M. J., Elliott, S. J. & Pinnington, R. J. 1994 Piezoelectric actuators and
sensors. In Proc. IUTAM Symp. on the Active Control of Vibration, pp. 263–274. Eindhoven:
International Union of Theoretical and Applied Mechanics.
Brennan, M. J., Elliott, S. J., Bonello, P. & Vincent, J. F. V. 2003 The ‘click’ mechanism in
dipteran flight: if it exists, then what effect does it have? J. Theor. Biol. 224, 205–213. (doi:10.
1016/S0022-5193(03)00158-9.)
Bonello, P., Brennan, M. J. & Elliott, S. J. 2004 The design of an adaptive tuned vibration
absorber with a variable shape stiffness element. ISVR Memorandum No. 941, ISVR, University
of Southampton, UK.
Bonello, P., Brennan, M. J. & Elliott, S. J. 2005 Vibration control using an adaptive tuned vibration
absorber with a variable curvature stiffness element. Smart Mater. Struct. 14, 1055–1065.
Carneal, J. P., Charette, F. & Fuller, C. R. 2004 Minimization of sound radiation from a plate
using adaptive tuned vibration absorbers. J. Sound Vib. 270, 781–792. (doi:10.1016/S0022-
460X(03)00257-8.)
Face International 1998 THUNDER TH-7R data sheet. Face International Corporation, Norfolk,
VA, USA.
Franchek, M. A., Ryan, M. W. & Bernhard, R. J. 1995 Adaptive-passive vibration control.
J. Sound Vib. 189, 565–585. (doi:10.1006/jsvi.1996.0037.)
Hong, D. P., Ryu, Y. S. 1985 Automatically controlled vibration absorber. US Patent No. 4935651.
Jones, R. M. 1975 Mechanics of composite materials. New York: Hemisphere.
Kidner, M. R. F. & Brennan, M. J. 1999 Improving the performance of a vibration neutraliser by
actively removing damping. J. Sound Vib. 221, 587–606. (doi:10.1006/jsvi.1998.2027.)
Kidner, M. R. F. & Brennan, M. J. 2002 Variable stiffness of a beam-like neutraliser under fuzzy
logic control. Trans. ASME J. Vib. Acoust. 124, 90–99. (doi:10.1115/1.1423634.)
Longbottom, C. J., Day, M. J., Rider, E. 1990 A self tuning vibration absorber. UK Patent No. GB
218957B.
Ormondroyd, J. & den Hartog, J. P. 1928 Theory of the dynamic absorber. Trans. ASME 50, 9–22.
Reissner, E. 1954 Note on the problem of vibrations of slightly curved bars. Trans. ASME J. Appl.
Mech. June, pp. 195–196.
Rustighi, E., Brennan, M. J. & Mace, B. R. 2005 A shape memory alloy tuned vibration absorber:
design and implementation. Smart Mater. Struct. 14, 19–28. (doi:10.1088/0964-1726/14/1/002.)
Ryder, G. H. 1986 Strength of materials 3rd edn. Basingstoke, UK: Macmillan.
Santulli, C., Patel, S. I., Jeronimidis, G., Davis, F. J. & Mitchell, G. R. 2002 Development of smart
actuators using N-isopropylacrylamide gels. In First World Congress on Biomimetics &
Artificial Muscles, pp. 1–9.
Srinivasan, A. V. & Michael McFarland, D. 2001 Smart structures: analysis and design.
Cambridge: Cambridge University Press.
Vincent, J. F. V. & Jeronimidis, G. 1991 The mechanical design of fossil plants. In Biomechanics
and evolution (ed. J. M. V. Rayner & R. J. Wootton), pp. 21–36. Cambridge: Cambridge
University Press.
von Flotow, A. H., Beard, A. H. & Bailey, D. 1994 Adaptive tuned vibration absorbers: tuning
laws, tracking agility, sizing and physical implementation. In Proc. Noise-Conf. 94, Florida,
pp. 81–101.
Walsh, P. L. & Lamancusa, J. S. 1992 A variable stiffness vibration absorber for the minimization
of transient vibrations. J. Sound Vib. 158, 195–211. (doi:10.1016/0022-460X(92)90045-Y.)

As this paper exceeds the maximum length normally permitted,


the authors have agreed to contribute to production costs.

Proc. R. Soc. A (2005)

Anda mungkin juga menyukai