You are on page 1of 13

A State-Switched Absorber Used

for Vibration Control of


Continuous Systems
A state-switched absorber (SSA) is a device that is capable of switching between discrete
stiffnesses; thus, it is able to instantaneously switch between resonance frequencies. The
Mark H. Holdhusen state-switched absorber is essentially a passive vibration absorber between switch
University of Wisconsin-Marathon County, events; however, at each switch event the SSA instantly “retunes” its natural frequency
518 S 7th Ave., and maintains that frequency until the next switch event. This paper considers the opti-
Wausau, WI 54401 mization of the state-switched absorber applied to a continuous vibrating system and
e-mail: mholdhus@uwc.edu details the experimental validation of these simulation results. A simulated annealing
optimization algorithm was utilized to optimize the state-switched absorber. For the most
Kenneth A. Cunefare part, the SSA performed only marginally better than a classical tuned vibration absorber
The George W. Woodruff School of Mechanical (TVA). However, for a select few cases considered, the SSA was able to reduce the kinetic
Engineering, energy of the plate to which it is attached by 12.9 dB over that of a classical tuned
The Georgia Institute of Technology, vibration absorber. The optimal SSA location on a clamped square plate was near the
Atlanta, GA 30332-0405 center of the plate for the vast majority of the forcing cases considered. To experimentally
e-mail: ken.cunefare@me.gatech.edu validate the simulation, a SSA was fabricated by employing magnetorheological elas-
tomers to achieve a stiffness change. For several two-force component excitations, sev-
eral tuning configurations of the SSA were applied and the kinetic energy of the system
was found and optimized. As with the majority of the optimization cases, the experiments
showed the SSA outperforming the TVA by only 2 dB. When comparing the observed
results to those found via simulation, the simulations accurately predicted the perfor-
mance of the SSA in the experiments. 关DOI: 10.1115/1.2748465兴

Keywords: state-switching, vibration absorbers, vibration control, optimization, simu-


lated annealing, magnetorheological elastomers

Introduction means to control vibrating systems. Davis et al. 关4兴 recognized


that piezoceramic devices could be considered as a controllable
This research considers the optimization and experimental vali-
variable stiffness element, due to the dependence of the piezoce-
dation of a state-switched absorber 共SSA兲 used to control the vi-
ramic’s stiffness on the output impedance. Clark 关5兴 and Richard
bration of continuous systems, specifically, a clamped plate. A
et al. 关6兴 have considered piezoelectric devices for state-switched
state-switched absorber is a dynamic device that is capable of vibration control. The switching criteria used by both Clark 关5兴
instantaneously switching between discrete stiffnesses and, thus, and Richard et al. 关6兴 allow for switching at maximum displace-
is capable of instantaneously switching between discrete reso- ment across the stiffness element to optimize the dissipation of
nance frequencies. The SSA is similar to a classical passive vibra- energy. However, switching at maximum strain may cause unde-
tion absorber, such as a tuned vibration absorber 共TVA兲 or a tuned sirable shocks in response due to a sudden release or addition of
vibration damper 共TVD兲. A TVA passively controls vibration of energy to the system 关7–10兴. State-switched systems are not, in
the base to which it is attached over a narrow frequency band general, stable 关11兴. Certain switching laws can lead to an energy
around the absorber’s resonance frequency 关1兴. A tuned vibration addition to the system, which can cause the system to become
damper can increase the effective bandwidth of the device by unstable 关11兴.
properly selecting the damping. However, there is trade-off in that The state-switched absorber concept was directly derived from
the TVD has decreased performance versus the TVA at the absorb- a device designed by Larson and Rogers 关7兴, Larson 关8兴, and
er’s resonance frequency 关1兴. The main difference between these Larson et al. 关9兴 for use in underwater transduction. The state-
classical absorbers and the SSA is that the resonance frequency of switched transducer has the same fundamental concept as the SSA
a classical absorber remains constant during operation, whereas in that it is capable of instantaneously switching between discrete
the SSA can switch between discrete resonance frequencies, thus stiffnesses. Switching between discrete stiffnesses changes the
increasing the effective bandwidth of the SSA. When a switch resonance frequency of the device, thus increasing the effective
event occurs, the state-switched absorber’s frequency changes in- bandwidth. The state-switched transducer uses piezoelectric mate-
stantaneously and remains at this new frequency until the next rial as the spring element in the system. The piezoelectric configu-
switch event. Therefore, between switch events the SSA is a pas- ration used by Larson and Rogers 关7兴, Larson 关8兴, and Larson et
sive device. The SSA has shown improved performance over clas- al. 关9兴 is similar to the approach used by Davis et al. 关4兴 and Davis
sical passive devices, especially when subjected to a multifre- and Lesieutre 关12兴, except the device of 关4,12兴 was implemented
quency excitation 关2,3兴. for vibration control, whereas the objective of 关7–9兴 was under-
State switching has been considered in previous research as a water transduction. The stiffness of a piezoelectric element is dif-
ferent when it is short circuited as compared to when it is open
circuited 关4,12兴. If the piezoelectric circuit can be controlled,
Contributed by the Technical Committee on Vibration and Sound of ASME for
publication in the JOURNAL OF VIBRATION AND ACOUSTICS. Manuscript received March
switching the stiffness of the piezoelectric material can occur as
8, 2006; final manuscript received May 4, 2007. Assoc. Editor William W. Clark. fast as the circuit can be changed. If the instantaneous change in
Paper presented at the SPIE Symposium on Smart Structures and Materials, 2004. stiffness occurs at zero strain, then no mechanical transients will

Journal of Vibration and Acoustics Copyright © 2007 by ASME OCTOBER 2007, Vol. 129 / 577

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
be introduced to the system. Switch events can potentially occur
every half cycle of oscillation of the state-switched device.
Piezoelectric elements have a relatively high stiffness, which
can lead to relatively large absorber masses for low frequencies of
interest. Also, piezoelectric materials should not be put into ten-
sion because they are fragile and have displacement amplitude
limitations. To make the device more robust and lower the ab-
sorber mass, magnetorheological 共MR兲 elastomers could be
implemented as a variable spring element 关13兴. An MR elastomer
is a polymer with fine iron particles suspended in it. The stiffness
of an MR elastomer depends on the magnetic flux across it 关13兴.
Since the MR elastomer is composed of rubber as opposed to
ceramic, the stiffness will be much lower than a piezoelectric Fig. 1 Model of state-switched dynamical device
material, thus resulting in a smaller mass for a similar resonance
frequency.
Cunefare et al. 关3兴 conducted introductory research on the state- This paper will continue with a brief discussion of the state-
switched absorber used for vibration control. Their work focused switching concept and switching rules. The basic equations of
on state-switching theory and theoretical simulations of simple motion used for the plate system considered by this research will
one-degree-of-freedom and two-degree-of-freedom systems. For then be outlined. The next section will explain the optimization
the single-degree-of-freedom system, an SSA with two discrete method employed and the optimization results. Then, the experi-
stiffnesses was compared to two passive tuned vibration absorbers mental setup will be detailed, including the absorber design, in-
共TVA兲 excited by an identical base motion. Cunefare et al. 关3兴 strumentation, and procedures. Results from the experiments are
found that the SSA dissipates more energy than the TVAs, illus- then outlined. Finally, some conclusions will be made concerning
trating a greater damping capability in the SSA. Also, as the spac- the research.
ing in excitation frequencies increased, the relative performance
of the SSA increased as compared to the TVAs. Therefore, the State-Switching Theory
SSA has a larger effective bandwidth as compared to the classical Figure 1 depicts a schematic of a state-switched absorber. The
TVA. SSA is a one-degree-of-freedom mass-spring-damper device
The second system investigated by Cunefare et al. 关3兴 was a whose spring can instantaneously switch between two discrete
simple two-degree-of-freedom system. A single two-state SSA stiffnesses. The arrow through the spring in Fig. 1 represents the
was compared to a single TVA attached to identical bases with switching capability of the stiffness element. At any time during
identical two-frequency component forcings. The SSA did as well operation, the SSA observes only one resonance frequency. The
as or outperformed the TVA for all combinations of excitation current state of the SSA is called the “on-line” state. Thus, the
frequencies investigated, including single frequency excitations. “off-line” state is the other potential state of the SSA. When a
As in the single-degree-of-freedom system, the comparative per- switch in state occurs the on-line stiffness becomes the off-line
formance of the SSA increases as the spacing between forcing stiffness and vice versa.
frequencies increases, illustrating a broader effective bandwidth. To avoid mechanical shocks in the system response, switching
The significance of the simulations is that the SSA has the poten- should only occur at zero relative displacement across the SSA.
tial for performance gains as compared to classical passive TVAs. When the relative displacement across the spring is zero, there is
Holdhusen 关14兴 also researched the SSA considering the role of no potential energy in the spring 共it is presumed that there is no
damping and its effect on the vibration control performance and static preload in the spring兲; thus, the energy before and after the
modeling of the state-switched absorber. The SSA was found to be switch event remains equal. Since the energy before the switch is
most effective at low damping values of the absorber. Also, the equal to the energy after the switch, thus, the system will remain
damping model used 共viscous or proportional兲 has no significant stable 关11,15兴. If a switch event from the upper frequency to the
effect on the final performance of the SSA as compared to a TVA. lower frequency occurs when the relative displacement is non-
Holdhusen 关14兴 also examined the experimental validation of zero, then the spring contains some potential energy that is instan-
the state-switched absorber. The state-switched absorber was taneously released, causing shocks in the response of the system.
implemented using a clutch and a collection of coil springs. The Zero strain across the spring is a necessary condition for
SSA showed performance gains experimentally as compared to switching to occur; however, the system needs not switch at every
classical tuned vibration absorbers. Also, the SSA was observed to occurrence of zero relative displacement. For the work previously
be effective over a larger bandwidth than that of a classical TVA. done by Cunefare et al. 关3兴 and Holdhusen 关14兴, a maximum work
Holdhusen and Cunefare 关15兴 investigated the maximum work extraction principle was implemented. To achieve a maximum
extraction switching rule, which maximizes the power dissipated

再 冎
work extraction in a two-state SSA, the switching rule is
by the absorber’s damper. They determined that since switching
has little affect on the power input to the system and all power kSSA1 ⇒ vbase共vSSA − vbase兲 ⬎ 0
next
kSSA = 共1兲
input to the system must be dissipated by the system’s dampers, kSSA2 ⇒ vbase共vSSA − vbase兲 ⬍ 0
maximizing the work extracted by the absorber’s damper mini- next
where kSSA is the stiffness to which to switch, kSSA1 is the lower
mizes the base motion. Also, the switching criteria allowing
of the two stiffnesses, kSSA2 is the larger of the two stiffnesses,
switching to occur only at zero relative displacement of the ab-
vbase is the velocity of the base at the point of attachment, and
sorber results in a stable system 关15兴.
The work at hand investigates the state-switched absorber ap- vSSA is the velocity of the SSA. It should be noted that the switch-
ing logic described by Eq. 共1兲 is also closely related to the “sky-
plied to a vibrating plate. Both the location of the absorber and the
hook” optimal damping concept 关16–18兴, except that the sky-hook
tuning frequencies of the absorber must be optimized to achieve
concept switches damping, whereas the SSA switches stiffness.
proper control of the system. This research optimizes these tuning
parameters for a state-switched absorber applied to a clamped
plate. The research at hand also experimentally validates the state- Basic Equations of Motion
switched absorber used for controlling the vibration of the Figure 2 depicts a rectangular plate clamped on all sides with a
clamped plate system. The performance of the SSA is compared to state-switched absorber attached to control vibrations. The equa-
that of an optimized classical tuned vibration absorber. tions of motion for such a system are of the form

578 / Vol. 129, OCTOBER 2007 Transactions of the ASME

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
The kinetic energy of the plate with absorbers attached is ex-
pressed as

冕冕
Lx Ly N+NA


1 1
T= m共x,y兲ẇ 共x,y兲dydx +
2
maq̇2a 共4兲
2 0 0
2 a=N+1
where NA is the number of attached absorbers, m共x , y兲 is the mass
per unit length of the plate, and ma are the masses of the attached
absorbers. The elements of the mass matrix are

冕冕
冦 冧
Lx Ly
m␸i␸ jdydx i, j 艋 N
0 0
mij = 共5兲
Fig. 2 State-switched absorber attached to a plate, subject to ma i, j ⬎ N
multiharmonic point forcing
0 i⫽j⬎N
The absorber mass is one-tenth the mass of the plate. The poten-
tial energy for the plate and attached absorbers may be expressed
Mq̈ + Cq̇ + Kiq = U 共2兲 as 关20兴

冕冕
where M is the mass matrix, C is the damping matrix, U is a Lx Ly
1
vector of forces, q is a vector of coordinates, and Ki is the stiffness V= D共x,y兲兵w2xx + w2yy + 2␯wxxwyy + 2共1 − ␯兲wxy
2
其dydx
2
matrix where the subscript i represents the current state of the 0 0
SSA. N+NA


The method of assumed modes is used to derive the equations 1
+ ka关qa − w共xa,y a兲兴2 共6兲
of motion 关19兴. The flexural displacement field on the plate is 2 a=N+1
represented as
N
where
w共x,y,t兲 = 兺 ␸ 共x,y兲q 共t兲 = ␸ q
i=1
i i
T
共3兲 D共x,y兲 =
E共x,y兲h共x,y兲3
12共1 − ␯2兲
共7兲

where N is the number of modes in the isolated plate, ␸ are basis E is the modulus of elasticity, ka are the absorbers’ stiffnesses, h is
functions, x and y define position on the plate, and q is a vector of the plate thickness, ␯ is Poisson’s ratio, and xa and y a define the
generalized coordinates. Position and time-dependency notation absorber’s position on the plate. The elements of the stiffness
will be dropped henceforth for simplicity. matrix are

冦 冧
冕冕
Lx Ly N+NA

0 0
L共␸i, ␸ j兲dydx + 兺
a=N+1
ka␸i共xa,y a兲␸ j共xa,y a兲 i, j 艋 N

kij = − ka␸i共xa,y a兲 i 艋 N, j ⬎ N 共8兲


− ka␸ j共xa,y a兲 j 艋 N,i ⬎ N
ka i=j⬎N
0 i, j ⬎ N and i ⫽ j

where
L共␸i, ␸ j兲 = D关␸xx,i␸xx,j + ␸yy,i␸yy,j
␸i共x,y兲 =
x
Lx
冉 冊 冉 冊 冉 冊 冉 冊
1−
x
Lx
sin
m␲x y
Lx L y
1−
y
Ly
sin
n␲ y
Ly
共10兲

+ 2␯␸xx,i␸yy,j + 2共1 − ␯兲␸xy,i␸xy,j兴 共9兲 where m and n are integer indices mapping to the subscript i. The
number of degrees of freedom in the plate is determined by the
There are two discrete stiffness matrices representing each of the
two potential stiffnesses between which the absorber can switch. maximum values of m and n,
The research at hand considers a square plate that has been DOFplate = mn 共11兲
clamped on all four sides. The basis function used to satisfy the
boundary conditions of such a plate is The elements of the mass matrix for a clamped plate become

冕 冕 冋 冉 冊 冉 冊册 冉 冊 冉 冊 冉 冊 冉 冊
冦 冧
Lx Ly
x x y y 2
m␲x n␲ y r␲x s␲ y
␳h 1− 1− sin sin sin sin dydx i, j 艋 DOFplate
0 0
Lx Lx L y Ly Lx Ly Lx Ly
mij = 共12兲
ma i, j ⬎ DOFplate
0 i ⫽ j ⬎ DOFplate
where m and n map to the subscript i and r and s map to the subscript j. The elements of the stiffness matrix are

Journal of Vibration and Acoustics OCTOBER 2007, Vol. 129 / 579

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
冕冕

冦 冧
冋 冉 冊 冉 冊册 冉 冊 冉 冊 冉 冊 冉 冊
Lx Ly
xa xa y a ya 2
m␲xa n␲ y a r␲xa s␲ y a
L共␸i, ␸ j兲dydx + ka 1− 1− sin sin sin sin i, j 艋 DOFplate
Lx Lx L y Ly Lx Ly Lx Ly
0 0

− ka
xa
Lx
冉 冊 冉 冊 冉 冊 冉 冊
1−
xa
Lx
sin
m␲xa
Lx
ya
Ly
1−
ya
Ly
sin
n␲ y a
Ly
i 艋 DOFplate, j ⬎ DOFplate
共13兲
冉 冊 冉 冊 冉 冊 冉 冊
kij = .
xa xa r␲xa ya ya s␲ y a
− ka 1− sin 1− sin j 艋 DOFplate,i ⬎ DOFplate
Lx Lx Lx Ly Ly Ly
ka i = j ⬎ DOFplate
0 i, j ⬎ DOFplate and i ⫽ j

As with the mass matrix, m and n map to the subscript i and r and with the excitation frequencies, are normalized by the natural fre-
s map to the subscript j. L is calculated using Eq. 共9兲, again using quency of the fundamental mode of the continuous system with no
proper mapping between indices. The double integrals in Eqs. 共12兲 attachment.
and 共13兲 are calculated numerically using the MATLAB function Optimization techniques that assume a continuous but not nec-
DBLQUAD. essarily convex objective function to find local maxima and
To determine the forcing vector, a point force on the plate lo- minima cannot be employed for the state-switching system. This
cated at xF = 0.25Lx and y F = 0.375Ly is used. The resulting forcing is due to the discontinuous nature of the kinetic energy objective
vector is function as a function of tuning parameters. A slight change in one

冉 冊 冉 冊 冉 冊
of the SSA parameters could change the response of the system.
1 3 45 ␲m 3␲n This change in response can cause switch events to occur at dif-
Ui = F␸i L x, L y = F sin sin 共14兲
4 8 1024 4 8 ferent instances and cause a significant change in the average
A proportional damping is used to model the damping in this kinetic energy of the system. Because of the discrete nature of
system. The damping matrix is defined as switching, the average kinetic energy objective function changes
discontinuously as a function of tuning parameters. Such discon-
C = ␣共M + Kavg兲 共15兲 tinuities make gradient-based optimizers infeasible because an in-
where ␣ is a proportionality constant and Kavg is the mean of the finite slope will cause a failure in the optimization.
two SSA stiffness matrices. For the work at hand, the proportion- To optimize the absorber’s attachment location and tuning fre-
ality constant is equal to 0.05. An average of the two SSA stiffness quencies, a simulated annealing optimization method was em-
matrices is used in Eq. 共15兲 because the stiffness matrix can ployed. Metropolis et al. 关21兴 developed a simple algorithm that
change during the response of the system and a switch in stiffness provides an efficient simulation of the annealing process of mate-
would cause damping switching, which is not the focus of the rials. Kirkpatrick et al. 关22兴 recognized that this simulated anneal-
research. ing algorithm could be utilized for optimization problems. Opti-
mization by simulated annealing is beneficial in that it allows for
the transition out of local optimum; therefore, the procedure need
State-Switched Absorber Optimization not get stuck in a nonglobal optimum. For the state-switching
Optimization Model. The response of the state-switched ab- problem at hand, simulated annealing is advantageous not only
sorber system described above is calculated using a linear time because it will not get stuck in local minimum, but also because it
invariant 共LTI兲 state-space method. This method can be used be- does not require the use of gradients.
cause the system is linear and time invariant between switch The simulated annealing optimization algorithm begins at an
events. When a switch event does occur, all the matrices are up- initial “temperature” with an initial guess of the optimization pa-
dated to reflect the new on-line state with initial conditions and rameters, which correspond to an objective function value. The
the LTI simulation continues until the next switch event. Cunefare procedure randomly selects new parameters in the neighborhood
et al. 关3兴 detail a more in-depth coverage of this simulation of the previous parameters. These new parameters also have a
method. corresponding objective function value. If the new value of the
The goal of this research is to compare the response of an objective function is less than the previous objective function
optimized SSA to that of an optimized TVA. The objective func- value, then the move is allowed and the new parameters are ac-
tion to be minimized is the time-averaged kinetic energy of the cepted as the current parameters, assuming a minimum objective
continuous base system. The kinetic energy of the system is de- function is desired. If the new value of the objective function is
fined as greater than the previous objective function value, then the move
away from the local minimum would be allowed if

冉 冊
1
T = 2 q̇TMq̇ 共16兲
f − f new
where q̇ are the generalized velocities corresponding to the plate exp ⬎ random共0,1兲 共17兲
c
and M is the mass matrix corresponding to the plate. The kinetic
energy described in Eq. 共16兲 is averaged over one repetition of the In Eq. 共17兲, f is the current objective function value, f new is the
system’s response. Because of the switching in the system, the objective function value corresponding to the newly determined
SSA can never truly be in steady state. However, a repetition in parameters, c is the simulated annealing temperature, and “ran-
the response of the system can be observed. This repetition period dom共0,1兲” is a number between zero and one, chosen using a
is the interval over which the kinetic energy is averaged. random number generator. Note that the use of Eq. 共17兲 allows for
There are four parameters that can be altered to achieve an movement to a higher objective value; thus, the procedure need
optimum performance depending on if the SSA is applied to a not get stuck in a local minimum. Whether or not a move occurs,
plate. Two parameters are the location of the absorber’s attach- a new set of parameters is chosen in the neighborhood of the
ment point on the vibrating body. The other two parameters are current parameters and the process is repeated. After this process
the upper and lower tuning frequencies of the SSA. Note that if has been repeated a predetermined number of iterations at the
the two SSA frequencies are equal, the absorber behaves as a same temperature, the temperature is lowered a small amount and
classical tuned vibration absorber. These tuning frequencies, along the steps are repeated at this new temperature. The algorithm con-

580 / Vol. 129, OCTOBER 2007 Transactions of the ASME

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 3 Kinetic energy ratio, in decibels, of an optimized SSA system to an
untreated plate as function of two forcing frequencies

tinues until the objective function remains unchanged over a num-


ber of consecutive temperatures. At this point, the system is “fro-
zen” and the optimum is found.
Er = 10 log 冉 冊
TSSA
Tplate
共18兲

Optimization Results. The simulated annealing optimization where TSSA is the plate kinetic energy with the optimized SSA
algorithm described above was used to optimize the tuning fre- attached and Tplate is the kinetic energy of a plate with no absorber
quencies and the attachment location of the state-switched ab- attached. When this value is less than zero, the SSA reduces vi-
sorber applied to a vibrating plate. The plate was subjected to a bration in the plate. All frequencies in Fig. 3 have been normal-
two-frequency component point excitation. A two-frequency com- ized by the natural frequency of the first mode of the plate. Figure
ponent excitation was chosen to correlate with the initial studies 3 shows that the SSA is effective at reducing the vibrations of the
of the SSA performed by Holdusen and Cunefare 关2兴, Cunefare et plate for the entire range of frequencies considered. For most of
al. 关3兴 and Holdusen 关14兴. Any machine having two separate ro- the cases considered, the SSA reduced plate vibrations by
tating parts would experience such an excitation. A real-world ⬃12 dB. The optimized SSA reduced the kinetic energy of the
example of such a machine is a helicopter with a main rotor ro- plate the most when the two normalized forcing frequencies were
tating at one frequency while the tail rotor rotates at a separate 1.06 and 3.72. For this forcing case, the SSA reduced the kinetic
frequency. Each excitation frequency ranged from just below the energy of the plate by 21.2 dB. Although the SSA showed excel-
natural frequency of the first plate mode to just above the natural lent performance at this forcing case, this was an extreme case and
frequency of the plate’s fifth mode. There are 16 forcing fre- the SSA only performed at this level in a few forcing cases
quency steps over the range resulting in 136 different forcing considered.
combinations. As with many optimization problems, this optimi- The relative performance of the optimized state-switched ab-
zation took several days to complete using MATLAB. The number sorber as compared to an optimized tuned vibration absorber is
of frequency steps was chosen so that the optimization could be shown in Fig. 4. The ratio of the plate kinetic energies 共in deci-
performed in a reasonable amount of time. For each combination bels兲 is plotted as a function of the two forcing frequencies, nor-
of excitation frequencies, the state-switched absorber’s attachment malized by the first mode’s frequency. For nearly all of the forcing
location and tuning frequencies were optimized by determining combinations considered, the ratio value is less than zero and the
which combination resulted in the lowest plate kinetic energy. optimized SSA outperforms the optimized TVA. However, the
This was done using the simulated annealing algorithm outlined SSA only showed ⬃3 dB improvement over the classical device
previously. Since there are two values that define the attachment for most of the cases considered. The best relative performance of
location on the two-dimensional plate and two tuning frequencies, the SSA occurs when the normalized excitation frequencies are
there are four parameters to be optimized in the state-switched 1.06 and 3.72, where the optimized SSA reduces the plate kinetic
absorber applied to a plate system. For comparison of the perfor- energy 12.9 dB more than the optimized TVA does. As stated
mance of the SSA to classical devices, a similar optimization was earlier, this forcing case is a rarity and the SSA only performs this
done to optimize the location and frequency for a tuned vibration well for a few cases considered. Most of the other cases resulted
absorber. in a improvement of around 3 dB over a TVA.
Figure 3 depicts the performance of the state-switched absorber As can be seen in Fig. 4, there are forcing frequency combina-
as compared to an untreated plate. The kinetic energy ratio of the tions where the optimized TVA narrowly outperforms the opti-
SSA system to the untreated system is shown as a function of both mized SSA. For these forcing cases, the simulated annealing al-
forcing frequencies. The energy ratio is quantified on a decibel gorithm finds a less than optimal local minimum in optimizing the
scale as described by SSA. For these forcing cases, the algorithm could have found both

Journal of Vibration and Acoustics OCTOBER 2007, Vol. 129 / 581

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 4 Kinetic energy ratio, in decibels, of an optimized SSA plate system
to an optimized TVA plate system as a function of two forcing frequencies

SSA tuning frequencies to be equal to the optimized TVA fre- the length of the plate in the respective directions, and the fre-
quency, thus causing the SSA to act as the optimum TVA. Al- quencies are normalized by the natural frequency of the first plate
though the simulated annealing algorithm allows for moving out mode. As can be seen from these Figs. 5 and 6, the attachment
of a local minimum, it does not guarantee that the global mini- point was to found to be near 0.5 in both directions for nearly the
mum will found. Thus, the simulated annealing algorithm can still entire range of forcing frequencies. This corresponds to an attach-
get stuck in a local minimum, as was the case for a few scattered ment point near the geometric center of the plate. The first mode
forcing cases. However, for the vast majority of forcing combina- shape of the plate has the largest amplitude of vibration at the
tions, the optimized SSA outperformed the optimized TVA. center of the plate. Therefore, the absorber is controlling this
Figures 5 and 6 define the attachment point of the optimized mode of vibration when attached at the midpoint of the plate. To
state-switched absorber on the plate for the range of two- achieve optimal control of the plate, the SSA is attached near the
frequency excitations considered. Since the plate has two dimen- center of the plate to control the fundamental vibration mode for
sions, two values are required to define the attachment location on nearly all of the forcing cases considered.
the plate. The locations defined by Figs. 5 and 6 are normalized by Figures 7 and 8 together define the optimal state-switched ab-

Fig. 5 Optimum SSA location in the x direction on the plate as function of


two forcing frequencies

582 / Vol. 129, OCTOBER 2007 Transactions of the ASME

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 6 Optimum SSA location in the y direction on the plate as function of
two forcing frequencies

sorber tuning frequencies for a range forcing frequency combina- between discrete frequencies would not occur. It can be seen from
tions. Figure 7 plots the mean of the two optimized SSA tuning Fig. 8 that several forcing combinations that result in a frequency
frequencies as a function of the two forcing frequencies. All fre- ratio of one. At these forcing cases, the simulated annealing algo-
quencies are normalized by the natural frequency of the funda- rithm is choosing the SSA frequencies to be equal, thus choosing
mental mode of the isolated plate. As can be seen from Fig. 7, the a TVA. When the optimization was repeated for the TVA, it chose
average SSA tuning frequency increases with increasing forcing the exact same absorber as was found by the SSA optimization
frequency. This result is as expected since the tuning frequencies resulting in equal plate kinetic energies. In other words, the opti-
should be tuned near the excitation frequencies being controlled. mization found the best SSA to be a TVA for these forcing cases.
Figure 8 plots the ratio of the upper SSA frequency to the lower On the other hand, there were a few forcing cases where the
SSA frequency as a function of the two-excitation frequencies. optimization converges to an SSA and when performed again for
Note there are very few forcing cases that resulted in an SSA a TVA converged to a kinetic energy lower that that of the opti-
frequency ratio at or very near one. With such a tuning, the SSA mized SSA. Therefore, we can deduce that sometimes the simu-
would represent a classical tuned vibration absorber as switching lated annealing algorithm did not find this global optimum, when

Fig. 7 Mean of two SSA tuning frequencies normalized by first mode of


plate as function of two forcing frequencies

Journal of Vibration and Acoustics OCTOBER 2007, Vol. 129 / 583

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 8 Ratio of optimum SSA tuning frequencies for a plate as function of
two forcing frequencies

it existed, for the SSA applied to a plate. The plate is a complex to a plate was only 1.24. Finally, there is no analytical correlation
system having four optimization parameters that lead to many between the forcing frequencies and the optimized tuning frequen-
opportunities for local minimums. Although the simulated anneal- cies. The combination of Figs. 7 and 8, which defines optimal
ing algorithm always finds a low plate kinetic energy, it is not tuning frequencies, shows no trend in the tuning frequencies as a
guaranteed to find the lowest plate kinetic energy. function of forcing frequencies. Therefore, there is no way to
Figure 8 also seems to show that for forcing frequencies ⬍2, approximate analytically what the optimized tuning frequencies
where good performance of the SSA occurred, the tuning fre- should be based on the forcing frequencies.
quency ratio is ⬎6 for most of that frequency range. MR elas- The sensitivity of the optimization algorithm must be examined
tomers have a potential tuning frequency ratio of 4 关23兴. There- to determine the robustness of the optimization results. The sen-
fore, good performance of the SSA applied to a plate can require sitivity is examined by slightly perturbing one of the tuning pa-
large spacing between the tuning frequencies, perhaps larger than rameters of the best performing state-switched absorber. As shown
what is physically realizable. However, the tuning frequency ratio in Fig. 9, the best relative SSA performance occurs when the
corresponding to the best relative performance of the SSA applied normalized forcing frequencies are 1.06 and 3.72 and the SSA

Fig. 9 Kinetic energy ratio of SSA to optimized TVA as a function of tuning


parameters near the best performing SSA

584 / Vol. 129, OCTOBER 2007 Transactions of the ASME

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 10 Kinetic energy ratio of SSA to optimized TVA as a function of tuning
parameters near average performing SSA

reduces the kinetic energy 12.9 dB more than a TVA. To deter- upper tuning frequency can change the SSA performance by
mine the sensitivity of this optimization, each SSA frequency and 2.5 dB. Changes in the lower tuning frequency had very little
absorber to plate mass ratio is perturbed within a range of 10% effect on the SSA performance near this specific optimal tuning
above and below the optimal value while holding all other tuning case. This “average” SSA’s performance is much less sensitive to
parameters constant. The ratio of the plate kinetic energy due to tuning parameters than that of a high-performance point, where a
the SSA to the kinetic energy of the optimal TVA for this forcing 1% change in a tuning parameter can change the plate kinetic
case is shown in Fig. 9 for the perturbation range. Each tuning energy by 13 dB. This means that an SSA designed in this region
parameter is normalized by its optimal value that resulted in the is more robust than one designed for a high-performance region.
best SSA performance in the optimization. As can be seen, there is As with the best performing SSA, the simulation results are not
about a 13 dB decrease in kinetic energy for a narrow range sensitive to perturbations in the mass ratio and the SSA perfor-
around each of the optimized frequencies. The bandwidth of the mance improves as the mass ratio increases.
optimal tuning values is ⬃2% for the lower tuning frequency, and Given that the SSA only had a 3 dB improvement over a TVA
⬃6% for the upper tuning frequencies. This means that perturbing for most of the forcing cases considered and the highly sensitive
the tuning parameters, even by a small amount, can lead to large nature of the system performance to small changes in parameters,
differences in the performance of the SSA for this specific forcing it is not certain whether the SSA is robust enough to take the place
and tuning case. The simulation results are not sensitive to small of classical devices in controlling the vibration of continuous sys-
perturbations in the mass ratio. As can be seen in Fig. 9, the plate tems. When the cost and complexity of the setup of the SSA is
kinetic energy gradually decreases as the mass ratio increases. considered, it is not obvious whether the small performance im-
Similar perturbation results can be found from all the sharps peaks provements of the SSA are worth the increased complexity as
seen in Fig. 4. compared to the classical passive devices.
The optimization results for the best performing SSA are very
sensitive to small perturbations in the tuning parameters. This Experimental Setup
means that while the tuning parameters yield excellent perfor-
mance, the likelihood that the optimization finds this narrow range Switched Absorber Design. Magnetorheological elastomers
of optimized tuning parameters is very small. Also, when fabri- 共MRE兲 were used to fabricate an absorber that can switch between
cating a physical SSA there may be difficulty in narrowing in on discrete stiffnesses. An MR elastomer is a rubber that has fine iron
this small band where the SSA performs its best. For certain forc- particles dispersed throughout. A magnetic field is applied during
ing cases, mistuning the optimal parameters, even by 1%, can the cure of the elastomer to align the iron particles in chains. The
cause the SSA to show no performance gains versus a classical stiffness of the cured MRE is dependent on the magnetic field
TVA. applied across the elastomer; the higher the magnetic flux, the
The sensitivity of the optimization was also investigated for an greater the stiffness. This stiffness’ dependence on the magnetic
SSA that performed 0.5 dB better than a TVA, a more typical flux was utilized in designing an SSA with variable discrete stiff-
performance gain found from the optimization than the 12.9 dB nesses. For more detail on the properties of MR elastomers, please
performance gain investigated above. Figure 10 shows the SSA- refer to the paper by Albanese and Cunefare. 关23兴
to-TVA energy ratio for perturbations around the optimal tuning Figure 11 illustrates the magnetorheological elastomer imple-
parameters for this “average” performing SSA. All values are nor- mentation of the state-switched absorber. This absorber design
malized and displayed similarly to those in Fig. 9. Although the was employed for the experiments to validate the SSA’s perfor-
results are not as sensitive as the best performing SSAs, the per- mance in reducing vibrations in continuous systems. The design
formance of an “average” performing SSA is also sensitive to consists of four MREs and an electromagnetic coil, along with
small perturbations in the tuning frequencies. A 4% change in the steel to close the magnetic circuit. The electromagnetic coil is

Journal of Vibration and Acoustics OCTOBER 2007, Vol. 129 / 585

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 11 MR elastomer implementation of a state-switched
absorber

simply a spool of magnet wire wrapped around an iron core. By Fig. 13 Frequency response of plate
running an electric current through the coiled wire, a magnetic
flux is created through the core of the coil. As seen in Fig. 11, this
flux continues into the vertical steel members and then through the of the state-switched absorber. Note that this range is different
MR elastomers. Since the MREs contain iron particles aligned in from than the range used on the plate as a new absorber was built
chains, the flux moves through the elastomer, thus causing an for the plate experiments. As can be seen, the potential switching
increase in stiffness as compared to when no flux exists. The frequency range covers a number of natural frequencies of the
magnetic circuit is completed by thin steel strips along the top and plate. The SSA will attempt to control these vibration modes of
bottom of the absorber. These strips are separated by aluminum the plate.
rods and are rigidly attached to the base to which the absorber is
attached. Aluminum is used as the supports so the magnetic flux Instrumentation. Figure 14 depicts the experimental setup of
path would consistently pass through the MR elastomers. The the plate system, showing all the sensors and actuators. An elec-
electromagnet and the larger steel members become the absorber trodynamic shaker delivers the force to the plate. An impedance
mass 共132 g兲, while the MR elastomers are the stiffness and head measures this force as well as the acceleration of the plate at
damping elements of the absorber. the excitation point. The state-switched absorber’s base is rigidly
Figure 12 plots the resonance frequencies of the state-switched attached to the plate, and a displacement probe is rigidly attached
absorber described above as a function of the electrical current to the base of the SSA. This displacement probe measures the
supplied to the electromagnet. As can be seen, there is an increase relative displacement between the SSA and the plate at the ab-
in resonance frequency with an increase in electrical current. The sorber attachment point.
frequency with no current applied is 48 Hz, whereas at 6 A, the The system is controlled using a digital signal processing sys-
resonance frequency is 98.9 Hz. Therefore, the SSA has a capa- tem. To implement the switching control algorithm defined by Eq.
bility to change its frequency by 106% and its stiffness by 325%. 共1兲, the displacement probe and the accelerometer with the same
plate attachment point as the SSA are used. The displacement
Plate Setup. A clamped plate was used to validate the SSA’s probe determines when the absorber has passed through a zero
performance, as the natural frequencies of the plate modes are strain condition. By differentiating the displacement signal, the
closer together than those of the plate modes; therefore, more than relative velocity, which is needed in Eq. 共1兲, is calculated. To find
one plate mode is within the switching limits of the physical state- the base velocity needed to determine when to switch, the signal
switched absorber. A plate was designed so that several vibration from the accelerometer with the same attachment point as the SSA
modes would be within the range of frequencies between which is integrated. To perform a switch, an electrical current is sent to
the SSA is capable of switching. A square plate clamped on all the electromagnetic coil using a bipolar operational power supply.
four sides was used for the experiments, similar to that used in the Figure 15 depicts the locations of all five accelerometers, the
numerical optimization. The plate used in the experiments was impedance head, and the various attachment locations of the ab-
fabricated using a 0.0312 in. 共0.8 mm兲 steel sheet clamped so that sorber that are distributed on the plate and used to measure the
each side was 18.25 in. 共46.4 cm兲 in length. These dimensions response. Each ai in Fig. 15 represents a location of an acceler-
resulted in a plate that had a mass of 1.35 kg and an absorber to ometer. These were placed based on either corresponding to an
base mass ratio of 0.1. The frequency response of the plate is SSA location or to avoid a nodal point on the plate for the modes
shown in Fig. 13. To determine this frequency response, the plate of interest. F and a4 correspond to the connection point of the
was excited at point to avoid a plate node and measured using impedance head, which measures the acceleration and force at the
several accelerometers distributed on the plate. The region be-
tween dashed lines in Fig. 13 is the range of potential frequencies

Fig. 12 SSA resonance frequencies as a function of electro-


magnet current Fig. 14 Experimental setup with sensors and actuators

586 / Vol. 129, OCTOBER 2007 Transactions of the ASME

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
optimal SSA and the optimal TVA were found to compare their
performances. Also, the absorber was placed at each of three at-
tachment locations for each forcing case. These attachment points
can be seen in Fig. 15. To determine how well the simulations
model the actual system, the SSA performance was also found
using the simulations used in the optimization study for each of
these same forcing and tuning cases considered in the
experiments.
For each forcing frequency combination investigated experi-
mentally, the shaker is enabled and allowed to excite the plate.
While the voltage amplitude sent to shaker is held constant, each
tuning combination is stepped through, acquiring 5 s of the plate’s
response for each frequency combination. This acquisition time
was found to be adequate by running some trials and observing
the time to pseudo-steady state and period of repetition in the
response. Pseudo-steady state was reached in ⬍1 s, and the period
Fig. 15 Plate layout including accelerometers, force trans- of repetition in the plate response was on the order of 0.5 s.
ducer, and SSA locations The kinetic energy of the plate is calculated for each tuning
combination from the acquired response of the accelerometers by

冤冥
vaccel1
point of excitation. The excitation point was also chosen to avoid
a node on the plate. The points labeled SSAi represent potential vaccel2
attachment points of the state-switched absorber. Only one ab- 2T vaccel3
sorber is attached at one of these points on the plate at any given = vTv = 关vaccel1 vaccel2 vaccel3 vaccel4 vaccel5 vaccel6兴
time. These attachment points represent a set of three potential M vaccel4
attachment points stepped through to experimentally validate the vaccel5
SSA simulations. vaccel6
Procedures. The plate considered in the experiments was sub- 共19兲
jected to a range of two-frequency component excitations. Table 1
where M is the mass matrix of the plate and vaccel is the velocity
shows the excitation frequencies stepped through for the plate.
at each accelerometer attachment point, calculated by integrating
There are six discrete forcing frequencies stepped through for the
each accelerometer’s signal. The root-mean-square 共rms兲 value of
plate resulting in 21 forcing combinations, including single fre-
quency component excitations. These sets of excitation and tuning the kinetic energy is calculated for the final 2.5 s of the response
frequencies were also used in numerical simulations to predict the and normalized by the rms force, measured from the impedance
performance of a SSA as compared to a TVA. head. Even though the kinetic energy is not averaged over a rep-
For experimental comparison between an SSA and a TVA, the etition interval, the maximum possible error averaging over the
best performing TVA and SSA must be found for each forcing final 2.5 s given a maximum repetition interval of 0.5 s is a neg-
case. The attachment location and both tuning frequencies that ligible 0.012 dB. This final energy metric can be written as
result in the best performance of each absorber should be found. Trms
To find the tuning frequencies that resulted in the best absorber Tmetric = 共20兲
Frms
performance, a direct search was employed by stepping through a
range of SSA resonance frequencies. Table 1 contains the poten- The kinetic energy is normalized by the force for accurate com-
tial SSA frequencies and the corresponding electrical current parison between tests. The kinetic energy metric is calculated for
needed to achieve each frequency. There are seven discrete fre- all the tunings where the resonance frequencies are equal. The
quencies that make up the whole set of potential absorber tuning tuning that results in the lowest kinetic energy is the best perform-
frequencies. Combining these potential states into groups of two ing TVA. The same is done for the tuning cases with unequal
tuning frequencies, including equivalent frequencies where the ab- frequencies to determine the best performing SSA. Comparison of
sorber acts as a TVA, results in 28 different absorber tuning fre- the best performing SSA and TVA is done for each forcing case in
quency combinations. For each forcing case considered, each of the experiments.
the 28 absorber tuning combinations was stepped through to de- For each forcing and tuning combination, simulations were also
termine which tuning parameters resulted in the best performance done to find the SSA performance. These simulations were per-
for both the SSA 共when the two tuning frequencies are not equal兲 formed similarly to those done by above. The SSA performance
and the TVA 共when both tuning frequencies are equal兲. Both the found through simulations was force normalized similarly to that
of the experiments, which is described in Eq. 共20兲. The predicted
SSA performance was then compared to the performance found
Table 1 SSA tuning frequencies and forcing frequencies for experimentally for each forcing and tuning case considered in the
the plate experiments experiments. This comparison is done by calculating the ratio of
the experimental performance of the SSA to the performance
SSA frequencies found through the simulations for each forcing and tuning combi-
nation. As a metric of how well the simulations predicted the
Current Frequency Forcing frequencies observed performance, the standard deviation of all these ratios
共A兲 共Hz兲 共Hz兲 for the plate is then calculated.
0 56.9 56
1 58.4 63
2 62.1 72 Experimental Results
3 68.1 81 One purpose of the experiments was to compare the perfor-
4 77.9 91
5 91.3 103 mance achieved using an SSA to the performance achieved using
6 102.1 a TVA. The experiments were also done to determine the how
well the simulations predicted the performance of the SSA. Before

Journal of Vibration and Acoustics OCTOBER 2007, Vol. 129 / 587

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 16 Experimental kinetic energy ratio, in decibels, of the Fig. 17 Experimental kinetic energy ratio, in decibels, of a
plate with SSA attached to untreated plate plate with SSA attached to a plate with TVA attached

was in the center of the plate. The mean and the standard devia-
comparing the performance of the SSA to that of a TVA and
tion from the mean of the kinetic energies were calculated for
comparing the experimental results to the simulations, the experi-
each of the ten trials. To quantify the repeatability of the experi-
mental plate response with an SSA attached is compared to the
mental SSA, the standard deviation was divided by the mean. This
experimental response of an untreated plate. Figure 16 depicts the
ratio was calculated to be 0.144, which corresponds to 0.59 dB.
experimental ratio of the kinetic energy found of the plate with a
This indicates that the standard deviation in the plate kinetic en-
state-switched absorber attached to the kinetic energy of an un-
ergy found from these experiments is 14.4% of the mean kinetic
treated plate as a function of excitation frequencies normalized by
energy. This result follows the repeatability results found for
the frequency of the fundamental plate mode. The ratio is plotted
MREs by Albanese and Cunefare 关23兴.
in Fig. 16 using a decibel scale defined as

冉 冊
To determine how well the simulations predicted the SSA per-
TSSA formance, simulations were performed for each of tuning and
Tratio = 10 log 共21兲 forcing cases considered in the plate experiments. For each test
Tuntreated
case, the force normalized plate kinetic energy was calculated
When this ratio is ⬍0 dB, the SSA reduces the kinetic energy of from the simulation and divided by the experimental kinetic en-
the vibrating plate. As can be seen from Fig. 16, the SSA reduces ergy, giving a performance ratio of the simulations to the experi-
the vibration of the plate for the entire range of forcing frequen- ments. When this ratio equals 1, the simulation exactly predicts
cies considered. The state-switched absorber achieves its greatest the experiment. Therefore, the standard deviation from a ratio of 1
vibration reduction of 14.8 dB when both normalized forcing fre- was calculated from all the cases considered in the simulations.
quencies are 2.14, which is in the vicinity of the plate’s natural This standard deviation from 1 was found to be 0.114, corre-
frequency near the middle of the potential SSA frequency range. sponding to 0.47 dB. Note, that the standard deviation from ex-
This maximum is for a specific forcing case and only occurred periment to experiment of 0.59 dB is larger than the 0.47 dB stan-
near this level a few times, over the rest of the cases considered dard deviation between the experiment and the simulation.
the SSA reduced plate vibrations by and average of ⬃6 dB. Therefore, the simulation predicts the observed SSA performance
The results of the experiments comparing the performance of to within the repeatability limits of experiment.
the state-switched absorber to the tuned vibration absorber perfor-
mance are depicted in Fig. 17. Figure 17 plots the observed ratio
of the plate kinetic energy with the state-switched absorber at- Conclusion
tached to the plate kinetic energy with the tuned vibration ab- First, the state-switched absorber was optimized to control a
sorber attached versus the two forcing frequencies, normalized by vibrating continuous system, specifically a clamped plate. The
the plate’s fundamental frequency. Again, when this ratio is SSA tuning frequencies and location on the plate were optimized
⬍0 dB, the SSA reduces plate vibration more than a TVA in the using a simulated annealing optimization algorithm, and its result-
experiments. Over the entire range of forcing frequencies consid- ing base system kinetic energy was compared to that of an opti-
ered, the ratio is ⬍0 dB, meaning that the SSA outperforms the mized tuned vibration absorber. The SSA performed as well as, or
TVA for each forcing case. The best relative performance of the better than, a classical TVA for the entire range of forcing frequen-
SSA as compared to the TVA occurs when the normalized excita- cies of the plate system. At its best, the SSA can reduce the kinetic
tion frequencies are both 1.88, where the SSA outperforms the energy of the plate to which it is attached by 12.9 dB over that of
TVA by 2 dB. Considering the complexity of the SSA setup as an optimized TVA. However, this high performance was one of a
compared to the passive TVA, it could be argued that these small few isolated cases. For most of the cases considered, the SSA only
improvements are not worth the added cost and complexity of the performed about 3 dB better than the TVA. Also, this is a simu-
SSA system. lation result and the performance of the SSA found by this simu-
To determine the repeatability of the plate experiments, a single lation is highly sensitive to small perturbations in the tuning pa-
forcing and tuning case was selected and the experiment was re- rameters causing potential difficulties in fabricating an SSA to
peated ten times. The specific case here was when the normalized achieve these large performance gains found by the optimization.
forcing frequencies were 1.88 and 2.14, the normalized SSA tun- Given the marginal improvements of the SSA, the highly sensitive
ing frequencies were 1.69 and 2.03, and the SSA attachment point nature of the performance with small changes in tuning param-

588 / Vol. 129, OCTOBER 2007 Transactions of the ASME

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
eters, and the increased cost and complexity to implement the ␣ ⫽ proportional damping constant
SSA, the SSA may not be robust enough and inexpensive enough ␸ ⫽ basis function
to adequately complete with the simpler and less-expensive pas- ␯ ⫽ Poisson’s ratio
sive TVA. ␳ ⫽ density
Also, experiments and simulations were performed to validate
the performance of the state-switched absorber used to control References
vibrating continuous systems, specifically a clamped plate. A 关1兴 Den Hartog, J. P., 1956, Mechanical Vibrations, Dover, New York.
state-switched absorber was designed using magnetorheological 关2兴 Holdhusen, M. H., and Cunefare, K. A., 2002, “Experimental Vibration Con-
elastomers 共MREs兲. MREs, which are made from rubber with iron trol of a Two-Degree of Freedom, State-Switched Absorber System,” 2002
ASME International Mechanical Engineering Congress and Exposition, New
particles suspended throughout, can change stiffness based on the Orleans, Vol. 71, pp. 421–427.
magnetic field applied across them. The SSA built for these ex- 关3兴 Cunefare, K. A., De Rosa, S., Sadegh, N., and Larson, G., 2000 “State-
periments had a maximum frequency change of 106%. The plate Switched Absorber for Semi-Active Structural Control,” J. Intell. Mater. Syst.
system were subjected to several two-frequency component exci- Struct., 11, pp. 300–310.
关4兴 Davis, C. L., Lesieutre, G. A., and Dosch, J., 1997 “A Tunable Electrically
tations. For each forcing case, an experimental direct search of a Shunted Piezoceramic Vibration Absorber,” SPIE Symposium on Smart Struc-
set of potential tuning frequencies and attachment locations was tures and Materials, San Diego, Vol. 3045, pp. 51–59.
done to find the best performing SSA as well as the best perform- 关5兴 Clark, W. W., 2000 “Vibration Control With State-Switched Piezoelectric Ma-
ing TVA. For the clamped plate experiments, the SSA outper- terials,” J. Intell. Mater. Syst. Struct., 11, pp. 263–271.
关6兴 Richard, C., Guyomar, D., Audigier, D., and Ching, G., 1999 “Semi-Passive
formed the TVA by 2 dB. Again, this SSA performance is a nar- Damping Using Continuous Switching of Piezoelectric Devices,” SPIE Sym-
row improvement over the classical tuned vibration absorber, posium on Smart Structures and Materials, Newport Beach, CA, Vol. 3672, pp.
which is much more straightforward to implement. Simulations of 104–111.
each experimental forcing and tuning case were performed to de- 关7兴 Larson, G. D., and Rogers, P. H., 1994 “State Switched Acoustic Source,” J.
Acoust. Soc. Am., 96, 3317共A兲.
termine how well the model predicted the experimental results. 关8兴 Larson, G. D., 1996 “The Analysis and Realization of a State Switched Acous-
The standard deviation between the SSA simulations and the ex- tic Transducer,” Ph.D. thesis, Georgia Institute of Technology, Atlanta, GA.
periments was found to be less than the standard deviation in the 关9兴 Larson, G. D., Rogers, P. H., and Munk, W., 1998 “State Switched Transduc-
repeatability of the experiments. Therefore, the numerical simula- ers: A New Approach to High-Power, Low-Frequency, Underwater Projec-
tors,” J. Acoust. Soc. Am., 103, pp. 1428–1441.
tions predicted results similar to those found in the experiments. 关10兴 Larson, G. D., and Cunefare, K. A., 2004, “Quarter-Cycle Switching Control-
for Shunt-Switched Dampers,” ASME J. Vibr. Acoust., 126, pp. 278–283.
关11兴 Kurdila, A. J., Clark, W. W., Wang, W., and McDaniel, D. E., 2002, “Stability
Nomenclature of a Class of Real-Time Switched Piezoelectric Shunts,” J. Intell. Mater. Syst.
c ⫽ simulated annealing temperature Struct., 13共10兲, pp. 107–116.
关12兴 Davis, C. L., and Lesieutre, G. A., 1995, “A Modal Strain Energy Approach to
C ⫽ damping matrix the Prediction of Resistively Shunted Piezoceramic Damping,” J. Sound Vib.,
DOF ⫽ degrees of freedom 184, pp. 129–139.
E ⫽ modulus of elasticity 关13兴 Ginder, J. M., Nichols, M. E., Elie, L. D., and Clark, S. M., 2000,
“Controllable-Stiffness Components Based on Magnetorheological Elas-
Er ⫽ energy ratio 共dB兲 tomers,” SPIE Symposium on Smart Structures and Materials, Newport Beach,
f ⫽ objective function value CA, pp. 418–425.
F ⫽ force amplitude 关14兴 Holdhusen, M., 2002, “Experimental Validation and the Effect of Damping on
h ⫽ plate thickness the State-Switched Absorber Used for Vibration Control,” M.S. thesis, Georgia
Institute of Technology, Atlanta, GA.
K ⫽ stiffness matrix 关15兴 Holdhusen, M. H., and Cunefare, K. A., 2004, “Investigation of the Two-State,
ka ⫽ absorber stiffness Maximum Work Extraction Switching Rule of a State-Switched Absorber for
kSSA1 ⫽ lower SSA frequency Vibration Control,” The International Symposium on Active Control of Sound
kSSA2 ⫽ upper SSA frequency and Vibration, Williamsburg, VA, Institute of Noise Control Engineering of the
USA, Washington DC, CD Proceedings.
L ⫽ length of plate side 关16兴 Bender, E. K., 1968, “Optimum Linear Preview Control With Application to
M ⫽ mass matrix Vehicle Suspension,” ASME J. Basic Eng., 90, pp. 213–221.
m ⫽ mass per unit length 关17兴 Karnopp, D., 1990, “Design Principles for Vibration Control Systems Using
Semi-Active Dampers,” ASME J. Dyn. Syst., Meas., Control, 112, pp. 448–
ma ⫽ absorber mass 455.
N ⫽ assumed modes 关18兴 Krasnicki, E. J., 1980 “Comparison of Analytical and Experimental Results for
NA ⫽ number of absorbers a Semi-Active Vibration Isolator,” The Shock and Vibration Bulletin, 50, pp.
Q ⫽ response coordinates 69–76.
关19兴 Meirovitch, L., 1986, Elements of Vibration Analysis, McGraw-Hill, New
t ⫽ time York.
T ⫽ kinetic energy 关20兴 Young, D., 1950, “Vibration of Rectangular Plates by the Ritz Method,”
U ⫽ force vector ASME J. Appl. Mech., 17, pp. 448–453.
关21兴 Metropolis, N., Rosenbluth, A., Rosenbluth, M., Teller, A., and Teller, E.,
vaccel ⫽ accelerometer velocity 1953 “Equation of State Calculations by Fast Computing Machines,” J. Chem.
vbase ⫽ base velocity at attachment point Phys., 21, pp. 1087–1092.
vSSA ⫽ absorber velocity 关22兴 Kirkpatrick, S., Gelatt, C. D., and Vecchi, M. P., 1983, “Optimization by Simu-
w ⫽ flexural displacement lated Annealing,” Science, 220, pp. 671–680.
关23兴 Albanese, A. M., and Cunefare, K. A., 2003, “Properties of a Magnetorheo-
x ⫽ horizontal plate coordinate logical Semiactive Vibration Absorber,” SPIE Symposium on Smart Structures
y ⫽ vertical plate coordinate and Materials, San Diego, CA, Vol. 5052, pp. 36–43.

Journal of Vibration and Acoustics OCTOBER 2007, Vol. 129 / 589

Downloaded 12 May 2010 to 115.248.99.121. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

You might also like