Anda di halaman 1dari 28

Analytical modeling of viscoelastic dampers for structural and

vibration control
q
S.W. Park
*
Georgia Institute of Technology, Atlanta, GA 30332, USA
Received 28 March 2000
Abstract
Dierent approaches to the mathematical modeling of viscoelastic dampers are addressed and their theoretical basis
and performance are compared. The standard mechanical model (SMM) comprising linear springs and dashpots is
shown to accurately describe the broad-band rheological behavior of common viscoelastic dampers and be more ef-
cient than other models such as the fractional derivative model and the modied power law. The SMM renders a
Prony series expression for the modulus and compliance functions in the time domain, and the remarkable mathe-
matical eciency associated with the exponential basis functions of a Prony series greatly facilitates model calibration
and interconversion. While cumbersome, nonlinear regression is usually required for other models, a simple collocation
or least-squares method can be used to t the SMM to available experimental data. The model allows viscoelastic
material functions to be readily determined either directly from the experimental data or through interconversion from
a function established in another domain. Numerical examples on two common viscoelastic dampers demonstrate the
advantages of the SMM over fractional derivative and power-law models. Detailed computational procedures for tting
and interconversion are discussed and illustrated. Published experimental data from a viscoelastic liquid damper and a
viscoelastic solid damper are used in the examples. 2001 Elsevier Science Ltd. All rights reserved.
Keywords: Viscoelastic damper; Rheological model; Standard mechanical model; Prony series; Fitting; Interconversion
1. Introduction
Protection of constructed facilities from damaging natural hazards has become an increasingly impor-
tant issue. Recurring destructive seismic events and hurricanes in the United States and elsewhere point to a
compelling need for the development of eective protective systems against such hazards. Various means
have been developed and implemented over the years to control excessive structural response to environ-
mental forces induced by earthquakes or winds. For example, in passive structural control, energy dissi-
pation devices are added to a structure so that a large portion of the input energy can be dissipated through
International Journal of Solids and Structures 38 (2001) 80658092
www.elsevier.com/locate/ijsolstr
q
Disclaimer: the views expressed in this article are those of the author and do not necessarily represent the views of the FHWA.
*
Now at U.S. Federal Highway Administration, Turner-Fairbank Highway Research Center, Oce of Infrastructure R&D, HRDI/
PSI, 6300 Georgetown Pike, McLean, VA 22101, USA.
E-mail address: sunwoo.park@fhwa.dot.gov (S.W. Park).
0020-7683/01/$ - see front matter 2001 Elsevier Science Ltd. All rights reserved.
PII: S0020- 7683( 01) 00026- 9
these devices, thereby reducing energy dissipation demand on the original structure. Such devices in-
clude metallic yield dampers, friction dampers, viscous or viscoelastic dampers, and tuned mass dampers
(Housner et al., 1997; Soong and Dargush, 1977).
Viscoelastic dampers have long been used in the control of vibration and noise in aerospace structures
and industrial machines (Kerwin, 1959; Ross et al., 1959; Jones, 1980; Torvik, 1980; Morgenthaler, 1987;
Ader et al., 1995). Similar applications have been undertaken for civil engineering structures. A pioneering
example is the 10,000 viscoelastic dampers installed in the twin towers of the World Trade Center in New
York in 1969 to mitigate the eects of wind loads (Mahmoodi et al., 1987). This was followed by a number
of other similar applications in the United States and abroad. The implementation of viscoelastic dampers
for seismic mitigation has been realized more recently (Zhang et al., 1989; Zhang and Soong, 1992; Chang
et al., 1993, 1995; Hanson, 1993; Bergman and Hanson, 1993; Tsai, 1993, 1994; Tsai and Lee, 1993a,b; Li
and Tsai, 1994; Samali and Kwok, 1995; Aprile et al., 1997; Hayes et al., 1999; Shukla and Datta, 1999;
Zou and Ou, 2000). Lately, electrorheological and magnetorheological uid dampers whose rheological
properties vary with applied electric or magnetic eld have received keen attention for their potential
applications in semi-active structural and vibration control (Gavin et al., 1996a,b; Dyke et al., 1996, 1998;
Makris, 1997; Sunakoda et al., 2000; Xu et al., 2000; Yang et al., 2000).
Analysis of a structure that incorporates viscoelastic dampers normally requires an analytical charac-
terization of the rheological behavior of the dampers. Dierent approaches to the analytical modeling of
the rheological behavior of a linear viscoelastic system are available in the literature. A classical approach
uses a mechanical model comprising a combination of linear springs and dashpots (Bland, 1960; Findley
et al., 1976; Ferry, 1980; Christensen, 1982; Tschoegl, 1989). The stressstrain relation for a linear viscoelastic
system represented by a springdashpot mechanical model is commonly expressed in a dierential operator
form, and the time-domain material functions derived from such a model is expressed by a series of de-
caying exponentials, often referred to as a Prony series. The model has been proved to be consistent with
the molecular theory (Rouse, 1953; Ferry et al., 1955) and the thermodynamic theory (Biot, 1954; Schapery,
1964). The mechanical model analogs and corresponding Prony series representations have long been used
to express the material functions of linear viscoelastic media, and the related topics including tting and
interconversion are well established; see Fung (1965), Findley et al. (1976), Ferry (1980), and Tschoegl
(1989) for a comprehensive treatment of mechanical model theories.
A modeling approach based on fractional calculus has also received considerable attention and been
used in characterizing the rheological behavior of linear viscoelastic systems by a number of authors (e.g.,
Gemant, 1938; Smit and de Vries, 1970; Bagley and Torvik, 1983a,b; Rogers, 1983; Koeller, 1984). This
approach uses the framework of a standard springdashpot mechanical model except that the regular
dierential operators are replaced by fractional-order dierential operators. The primary motivation for the
use of fractional derivatives comes from their ability to describe the broad-band behavior of many vi-
scoelastic materials with a small number of parameters. Other widely used phenomenological models for
linear viscoelasticity include dierent forms of power laws (Schapery, 1974). In particular, the so-called
modied power law (MPL) (Williams, 1964), derived from the phenomenology of polymers, often provides
an excellent representation of the broad-band relaxation or creep behavior of amorphous polymers above
their glass transition temperature. Numerous other mathematical models for linear viscoelasticity are also
available (e.g., Tschoegl, 1989).
A review of the literature indicates that the fractional derivative model (FDM) has predominantly been
used for viscoelastic dampers (Koh and Kelly, 1990; Makris, 1991; Makris and Constantinou, 1991, 1992,
1993; Tsai and Lee, 1993a,b; Makris et al., 1993a,b, 1995; Aprile et al., 1997). For example, Koh and Kelly
(1990) modeled elastomeric bearings using a fractional-order Kelvin model and observed the superiority
of its performance to that of the standard Kelvin model. Makris and Constantinou (1991) modeled a
viscoelastic uid damper using a fractional-order Maxwell model and reported its advantage over the
standard Maxwell model in describing the viscoelastic material functions over a broad range of frequency.
8066 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
A MPL was also used to characterize the relaxation function of a viscoelastic solid damper (Shen and
Soong, 1995). This MPL, with a small number of parameters, was found to describe the broad-band be-
havior of the damper quite well both in the time domain and frequency domain. No particular report has
been identied that addresses the use of the standard mechanical model (SMM) for viscoelastic dampers.
Although FDM is capable of characterizing a broad-band viscoelastic behavior with a small number of
model constants, the complex mathematical expressions of material functions and associated cumbersome
numerical operations make the model less ecient for a routine implementation. Determination of model
constants from experimental data normally requires a dicult, nonlinear regression procedure. In addition,
interconversion between material functions is not always feasible with FDM. The use of a small number of
constants in FDM signies a limited degree of freedom associated with the model and consequently the
model often results in a crude representation of the given data. It is known that a power-law model is
intrinsically linked to FDM (Bagley, 1989), and, therefore, a power-law model has limitations similar to
those of FDM.
In contrast, SMM, due to its sound physical basis and remarkable computational eciency, lends itself
to a better alternative to other prevailing models in the characterization of viscoelastic dampers. Although
each term in a Prony representation can depict only a narrow-band behavior, the series as a whole can
describe a broad-band behavior very accurately. The ndings reported by some investigators (e.g., Koh and
Kelly, 1990; Makris and Constantinou, 1991) that the standard Voigt or Maxwell model does not ade-
quately describe the rheological behavior of viscoelastic dampers can be explained by the narrow-band
representation capabilities of these simple models. Although useful for a conceptual illustration of a vi-
scoelastic phenomenon, these simple models are not adequate for characterization of the broad-band be-
havior of real viscoelastic media (Tschoegl, 1989). Instead, a model composed of multiple Voigt or Maxwell
elements, i.e., the generalized Voigt or generalized Maxwell model, can be used to characterize the broad-
band rheological behavior of linear viscoelastic media.
In Section 2, the linear viscoelastic stressstrain relationship and dierent approaches to the modeling of
linear viscoelasticity are reviewed. In Section 3, characterization of the rheological behavior of viscoelastic
dampers using SMM is presented. Analytical representation of linear viscoelastic material functions in
dierent domains and the methods of tting and interconversion are discussed. Two numerical examples,
one on a viscoelastic liquid damper and the other on a viscoelastic solid damper, are presented in Sections 4
and 5, respectively. The examples illustrate the detailed procedures for tting and interconversion, and the
performance of SMM is compared with that of other models including FDM and MPL. A further dis-
cussion on the theoretical basis for dierent models discussed in the text are provided in Section 6.
2. Linear viscoelasticity and mathematical models
2.1. The standard mechanical model
The uniaxial, isothermal stressstrain equation for a nonaging, linear viscoelastic material can be rep-
resented by the following Boltzmann superposition integral:
r(t) =
_
t
0
E(t s)
de(s)
ds
ds (1)
where E(t) is the relaxation modulus, and E(t) = e(t) = 0 for < t < 0. Eq. (1) follows from the memory
hypothesis, smoothness assumptions and mathematical representation theorem (Christensen, 1982). For a
thermorheologically simple material (Morland and Lee, 1960), the stressstrain equation under transient
temperatures (or nonisothermal condition) can also be expressed by Eq. (1) but with the time variable, s,
replaced with the so-called reduced time dened as n =
_
t
0
ds=a
T
where a
T
is a function of temperature
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8067
called the timetemperature shift factor. Since Eq. (1) is founded on general principles, the equation is valid
for any linear viscoelastic material irrespective of the model employed to express the material function, E(t).
A number of phenomenological models of the behavior of linear viscoelastic materials are available in the
literature (Ferry, 1980; Tschoegl, 1989).
A classical approach to the modeling of linear viscoelastic behavior employs a mechanical model
composed of linear springs and dashpots, and the stressstrain equation for such a model involves standard
(or ordinary) dierential operators. A general form of the stressstrain equation in dierential operators is
given by (Fung, 1965)

M
m=0
a
m
d
m
r
dt
m
=

N
n=0
b
n
d
n
e
dt
n
(2)
where a
m
and b
n
are constants, and d
m
( )=dt
m
denotes the mth-order time derivative of the function ( ).
Mechanical models with dierent arrangements of springs and dashpots render dierent mechanical in-
terpretations of the constants a
m
and b
n
in Eq. (2).
The generalized Maxwell model, widely used to characterize the modulus functions of linear viscoelastic
media, consists of a spring and m Maxwell units connected in parallel as illustrated in Fig. 1(a). A series
combination of a spring and a dashpot constitutes a Maxwell unit. The relaxation modulus derived from
the generalized Maxwell model is given by (Tschoegl, 1989)
E(t) = E
e

m
i=1
E
i
e
t=q
i
(3)
where E
e
, E
i
and q
i
are all positive constants representing the equilibrium modulus, relaxation strengths and
relaxation times, respectively; the relaxation time of the ith Maxwell unit is dened by q
i
= g
i
=E
i
where g
i
is
the viscosity of the unit. A typical term under the summation symbol in Eq. (3) represents the relaxation
modulus of the ith Maxwell unit. The series expression in Eq. (3) is often referred to as a Prony or Dirichlet
series.
The specialized forms of the dierential operator equation, Eq. (2), for some common mechanical
models including the generalized Maxwell model and the generalized Voigt model are given by Findley
et al. (1976). Note that Eq. (3) is derived from a relation between the general stressstrain equation, Eq. (1)
Fig. 1. SMMs; (a) generalized Maxwell model, (b) generalized Voigt model.
8068 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
and the stressstrain equation in a dierential-operator form corresponding to the generalized Maxwell
model. The use of a mechanical model not only leads to an explicit form of the material function such as
Eq. (3) but also makes it possible to relate thermodynamic and molecular parameters to measured time- or
frequency-dependent mechanical properties (Ferry, 1980). The SMM (that invokes standard dierential
operators in its mathematical representation) has long been accepted as an accurate, ecient tool for the
characterization of the viscoelastic behavior of many polymers and polymeric composites.
2.2. The fractional derivative model
In an eort to further generalize the stressstrain equation in a dierential operator form, Eq. (2), a
number of authors (e.g., Gemant, 1938; Smit and de Vries, 1970; Bagley and Torvik, 1983a,b; Rogers, 1983;
Koeller, 1984) have applied a notion of the fractional derivative to the equation in such a way that

M
m=0
a
m
d
p
m
r
dt
p
m
=

N
n=0
b
n
d
q
n
e
dt
q
n
(4)
where p
m
and q
n
are real constants with 0 6p
m
, q
n
61. The ordinary time derivatives acting on the time-
dependent stress and strain elds in Eq. (2) are now replaced with corresponding fractional-order time
derivatives. A fractional time-derivative of order a is dened, in an integral form, as (Bagley and Torvik,
1983a)
d
a
f (t)
dt
a
=
1
C(1 a)
d
dt
_
t
0
f (s)
(t s)
a
ds
_ _
(5)
where 0 < a < 1 and C() denotes the gamma function.
Although a summation of multiple terms with dierent fractional orders are implied on both sides of
Eq. (4), only a few terms are used in practice. For example, a model with M = N = 1, p
0
= q
0
= 0 and
p
1
= q
1
= a in Eq. (4) is often used to describe the stressstrain behavior of a class of viscoelastic materials,
i.e.,
r b
d
a
r
dt
a
= E
0
e E
1
d
a
e
dt
a
(6)
where b, E
0
and E
1
are constants and 0 6a 61. An explicit form of the relaxation modulus E(t) can be
derived from Eqs. (1) and (6) but its expression would be much more involved than Eq. (3) derived from
SMM.
2.3. Power-law representations
Dierent forms of power law have long been used to describe the relaxation and creep behaviors of linear
viscoelastic materials (e.g, Nutting, 1921). In particular, a power law of the following form is widely used
for its simplicity:
E(t) = E
e
E
1
t
n
(7)
where E
e
, E
1
and n are positive constants; E
e
here denotes the equilibrium (or rubbery) modulus and has the
same physical meaning as that of E
e
in Eq. (3).
Although Eq. (7) describes the relaxation behavior in the rubbery and transition regions well, it does not
provide a good representation for the glassy behavior of the material because Eq. (7) renders unbounded
values at short times. This shortcoming has prompted an introduction of the so-called MPL of the fol-
lowing form (Williams, 1964):
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8069
E(t) = E
e

E
g
E
e
(1 t=q)
n
(8)
where E
e
, E
g
, q and n are all positive constants which represent, respectively, the equilibrium modulus,
glassy modulus, relaxation time and power. With its broad-band representation characteristic, Eq. (8) can
reasonably depict the glassy and rubbery plateau behavior as well as the transition-region behavior of a
large class of polymers and polymer-based composites. Note that when t=q 1, Eq. (8) reduces to Eq. (7).
It is informative to note that the FDM and power laws are closely interrelated (Bagley, 1989; Tschoegl,
1989). For example, Bagley (1989) matched E(t) derived from Eq. (6) with that of Eq. (8) and found the
following asymptotic relations between the two groups of model parameters:
E
0
= E
e
;
E
1
b
= E
g
; a = n;
b
C(1 a)
= q
n
(9)
However, the two models are not completely equivalent. The discrepancy lies primarily in their short-time
relaxation behaviors. Numerous other phenomenological models for the behavior of linear viscoelastic
materials are available in the literature (e.g., Tschoegl, 1989).
2.4. Linear viscoelastic material functions in dierent domains
It is well known that all linear viscoelastic material functions are mathematically equivalent for a given
mode of loading. Each function contains essentially the same rheological information of the material.
However, depending on the nature of the input excitation, a certain material function can be used ad-
vantageously over others in computing the response of a viscoelastic medium. For example, a modulus
function can be used more conveniently when strain is specied as the input, and a creep function for
problems with a stress input. Similarly, a time-dependent or frequency-dependent material function may be
used advantageously when the input is transient or steady-state harmonic, respectively. Further, in solving
viscoelastic boundary-value problems following the Laplace transform-based elastic-viscoelastic corre-
spondence principle, one needs to deal with material functions dened in the Laplace-transform domain.
For later references, the linear viscoelastic material functions in dierent domains are briey reviewed in
Appendix A.
3. Characterization of viscoelastic damers using the standard mechanical model
Characterization of the rheological behavior of a viscoelastic damper requires the knowledge of the
geometry and material properties of each individual component constituting the damper unit (or system). A
constitutive relation for the damper is determined by relating the macroscopic response of the unit with the
applied excitation. A rigorous treatment entails the solution of a boundary value problem dealing with the
damper unit as a whole (e.g., Makris et al., 1995). However, more practically, a macroscopic (or eective)
material function of the damper system can be obtained from a physical experiment in which the system
input and output values are measured and related (e.g., Makris and Constantinou, 1991). A direct ex-
perimental characterization is simple and straightforward and does not require an explicit account of the
individual components of the damper system. The rheological behavior of a damper may be described
analytically using a mathematical model. The SMM is adopted in this paper as a prefered tool for char-
acterization of the rheological behavior of viscoelastic dampers. The details of the modeling procedure are
illustrated through the use of experimental data from common viscoelastic dampers. First, analytical ex-
pressions for the general linear viscoelastic material functions based on the SMM are discussed. Then, the
procedure for model tting and the issue on interconversion between material functions are addressed.
8070 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
3.1. Linear viscoelastic material functions
The Prony series representation of the relaxation modulus derived from the generalized Maxwell model
was given in Eq. (3). Similarly, the creep compliance can be conveniently characterized by the generalized
Voigt model that comprises a spring, a dashpot and n Voigt units connected in series, see Fig. 1(b),
D(t) = D
g

t
g
0

n
j=1
D
j
(1 e
t=s
j
) (10)
where D
g
, g
0
, D
j
and s
j
are all positive constants denoting the glassy compliance, zero shear-rate viscosity,
retardation strengths and retardation times, respectively; the retardation time of the jth Voigt unit is dened
by s
j
= D
j
g
j
where g
j
is the viscosity of the unit. For viscoelastic solids, g
0
and thus the second term in
Eq. (10) vanishes. For viscoelastic uids, g
0
is nite. A typical term under the summation symbol in Eq. (10)
represents the creep compliance of the jth Voigt unit consisting of a parallel combination of a spring and a
dashpot.
Now the time-domain material functions, Eqs. (3) and (10), can be substituted into Eqs. (A.7)(A.10)
and (A.16), (A.17) to obtain the corresponding material functions in the frequency and Laplace-transform
domains as follows (Tschoegl, 1989):
E
/
(x) = E
e

m
i=1
x
2
q
2
i
E
i
x
2
q
2
i
1
(11)
E
//
(x) =

m
i=1
xq
i
E
i
x
2
q
2
i
1
(12)
D
/
(x) = D
g

n
j=1
D
j
x
2
s
2
j
1
(13)
D
//
(x) =
1
g
0
x

n
j=1
xs
j
D
j
x
2
s
2
j
1
(14)
~
E(s) = E
e

m
i=1
sq
i
E
i
sq
i
1
(15)
~
D(s) = D
g

1
g
0
s

n
j=1
D
j
ss
j
1
(16)
where E
/
, E
//
, D
/
and D
//
are commonly referred to as the storage modulus, loss modulus, storage compliance
and loss compliance, respectively, and
~
E = s

E and
~
D = s

D are often called the operational modulus and


compliance where

E and

D denote the Laplace transforms of E(t) and D(t), respectively. The symbols x and
s denote the circular frequency and the Laplace transform parameter, respectively. Eqs. (3) and (10)(16)
indicate that once a material function (either modulus or compliance) is determined in a particular domain,
the corresponding material functions in other domains are automatically established in terms of the same
model constants. The compact, closed-form expressions for the material functions in the frequency and
Laplace-transform domains are due to the amenable operational properties associated with the exponential
basis functions in the series, Eqs. (3) and (10).
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8071
3.2. Fitting of material functions
The constants in the series representation of a material function can be evaluated by tting the ex-
pression to available experimental data. Various methods of tting have been introduced by others. For
example, in the collocation method, Eq. (3) is equated to the data measured at m dierent times, and the m
unknowns, E
i
(i = 1; . . . ; m), are found by solving the resulting system of m linear algebraic equations. In
the least squares method, the equation is equated to the data at more than m sampling points and the
resulting over-determined system is solved by minimizing the square errors. Schapery (1961) illustrated the
collocation method of tting using the relaxation modulus data from PMMA and shear storage compliance
data from polyisobutylene. Cost and Becker (1970) used the so-called multidata method (based in a least
squares scheme) to determine the model constants within the Laplace-transform domain. In both methods,
one faces 2m unknowns including E
i
and q
i
(i = 1; . . . ; m) in general, which would lead to a system of 2m
nonlinear equations. However, the relaxation times q
i
are usually specied a priori from experience and
only E
i
are determined by solving the resulting system of m linear equations. Selection of time constants
is discussed and illustrated in Sections 4 and 5. The equilibrium modulus, E
e
, is usually estimated by in-
specting the long-time behavior of the relaxation modulus data. It is to be noted that the model parameters
in a modulus function may equally be determined by tting Eq. (11) or Eq. (12) to the frequency-domain
experimental data, if available. Similarly, model parameters in a compliance function may be found by
tting Eqs. (10), (13) or (14) to available experimental data in the time or frequency domain.
The issue of tting the SMM to experimental data has been extensively discussed in the literature.
Recently, a number of researchers have proposed various techniques for improvement of traditional tting
methods. For instance, Emri and Tschoegl (1993, 1994, 1995) and Tschoegl and Emri (1992, 1993) used
only well-dened subsets of the experimental data to enhance the quality of a t, Kaschta and Schwarzl
(1994a,b) proposed a method that ensures positive coecients through an interactive adjustment of re-
laxation or retardation times, and Baumgaertel and Winter (1989) employed a nonlinear regression tech-
nique in which the coecients, time constants, and the number of terms in the series are all variable. Mead
(1994) used a constrained linear regression with regularization, Honerkamp and Weese (1989) and Elster
et al. (1991) applied the so-called Tikhonov regularization techniques, and Elster and Honerkamp (1991)
used the modied maximum entropy method to determine discrete viscoelastic spectra from rheological
measurements.
3.3. Interconversion between material functions
A linear viscoelastic material function can be converted into other equivalent material functions through
appropriate mathematical operations. Interconversion may be required for dierent reasons. The response
of a medium under a certain excitation condition inaccessible to direct experiment may be predicted from
measurements under other readily realizable conditions. For example, it is often dicult to subject sti
materials to a constant-strain, relaxation test because of the requirement of a robust testing device. How-
ever, a constant-stress, creep test is relatively easy to carry out on these materials. In this case, the relaxation
modulus can be determined from the measured creep compliance through an interconversion between the
relaxation modulus and creep compliance. Similarly, time-domain material functions can be obtained
through the conversion of corresponding frequency-domain material functions measured from steady-state
harmonic tests which usually yield more accurate information than quasistatic tests.
Numerous interconversion methods, either exact or approximate, have been proposed. Hopkins and
Hamming (1957) and Kno and Hopkins (1972) presented numerical interconversion techniques based
on the integral relationship between the relaxation modulus and creep compliance similar to Eq. (A.2),
Baumgaertel and Winter (1989) demonstrated an analytical conversion from the relaxation modulus to the
creep compliance using their interrelationship in the Laplace transform domain, and Mead (1994) presented
8072 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
a technique based on a constrained linear regression and used the technique to determine the relaxation
modulus from a set of storage and loss modulus data. Ramkumar et al. (1997) proposed a regularization
method that employs the quadratic programming originally developed for solving ill-posed Fredholm in-
tegral equations, and demonstrated the eectiveness of the method through the determination of the re-
laxation spectrum from steady-state experimental data. Park and Schapery (1999) presented a numerical
method of interconversion between linear viscoelastic material functions based on a Prony series repre-
sentation, and tested its eectiveness using experimental data from selected polymeric materials. Their
method is applicable to interconversion between modulus and compliance functions in time, frequency, and
Laplace transform domains. An approximate, analytical method of interconversion was presented by
Schapery and Park (1999). For a comprehensive treatment of the subject, the reader is referred to the
treatises by Schwarzl and Struik (1967), Ferry (1980), and Tschoegl (1989).
4. Numerical example 1 a viscoelastic liquid damper
The methods of constitutive modeling of viscoelastic dampers discussed above will now be illustrated
through two specic examples. The rst example concerns a viscoelastic liquid damper and the second a
viscoelastic solid damper. In each example, the performance of the SMM and a comparable model such as
the FDM or MPL is discussed and contrasted.
Viscoelastic liquids possess excellent energy dissipation characteristics and are widely used in dierent
types of energy dissipation device (Harris and Crede, 1976). A dashpot is a classical example of this kind.
Viscoelastic liquid dampers consisting of a cylindrical piston immersed in a viscoelastic uid (Schwahn and
Delinic, 1988) and viscous damping walls constructed of a at-plate piston immersed in a viscoelastic uid
(Arima et al., 1988) are commonly used to reduce vibration or isolate structures from seismic or wind
disturbances. Fig. 2(a) shows a viscous damper manufactured by GERB vibration control (GERB, 1986)
and used by Makris and Constantinou (1991) among others. The cylindrical pot is lled with silicon gel, a
highly viscous substance. The rheological characteristic of the damper depends on the viscoelastic prop-
erties of the uid and the geometric details of the damper unit. Fig. 2(b) displays typical forcedisplacement
hysteresis loops measured from vertical piston motion at the frequency of 2 Hz and at a room temperature
(Makris, 1991). Dampers of this type were used by Schwahn and Delinic (1988) for vibration control of
Fig. 2. A viscoelastic liquid damper; (a) geometry (Makris and Constantinou, 1991), (b) forcedisplacement hysteresis loops for vertical
motion at f = 2 Hz and T ~ 25C (Makris, 1991).
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8073
piping networks and by Humann (1985) for seismic base isolation of structures. The experimental data
presented by Makris (1991) and Makris and Constantinou (1991) for a viscoelastic liquid damper (or
simply, a viscous damper) are used here to illustrate how the data can be represented by analytical models
discussed above, and how the system response (or kernel) functions can be interconverted between dierent
analysis domains and loading modes. The term, viscous damper, is frequently used to refer to a viscoelastic
liquid (or uid) damper in the literature; however, it is to be noted that a viscoelastic uid possesses
elasticity as well as viscosity, while a viscous uid possesses only viscosity. The terms, viscous damper and
viscoelastic liquid damper, will be used interchangeably in this paper for convenience.
4.1. K
/
(x) and K
//
(x) by the fractional derivative model
The experimental data for the storage and loss stiness of a viscous damper (Fig. 2) in the longitudinal
motion of the piston are shown in Fig. 3(a) and (b), respectively, together with their analytical ts. Here,
stiness rather than modulus is used since a forcedisplacement (rather than stressstrain) relation is dis-
cussed. The experimental data presented in a tabular form by Makris (1991) are used. The SMM and the
FDM are compared here in their analytical representation of the experimental data. Makris (1991) and
Makris and Constantinou (1991), using FDM, obtained the following forcedisplacement relationship for
the longitudinal motion of the piston:
P k
d
r
P
dt
r
= C
0
d
q
u
dt
q
(17)
where P and u are the axial force and displacement, respectively, and the damper parameters C
0
, k, r and q
represent the zero-frequency damping coecient, relaxation time, and orders of fractional derivative, re-
spectively. When r = q = 1, the model reduces to the classical Maxwell model with k being the relaxation
time and C
0
the viscosity. Eq. (17) is slightly dierent from Eq. (6) besides the fact that r and e are replaced
with P and u. Taking the Fourier transform of Eq. (17) and rearranging terms, one nds expressions
parallel to Eqs. (A.11) and (A.5),
^
P(x) = K
+
(ix)^ u(x) (18)
K
+
(ix) = K
/
(x) iK
//
(x) (19)
where
^
P and ^ u denote the Fourier transforms of P and u, respectively, and K
+
is the complex stiness whose
real and imaginary components, K
/
and K
//
, are the storage and loss stiness given by (Makris and Con-
stantinou, 1991)
K
/
(x) =
C
0
x
q
cos(
pq
2
)[1 kx
r
cos(
pr
2
)[ C
0
kx
qr
sin(
pr
2
) sin(
pq
2
)
1 k
2
x
2r
2kx
r
cos(
pr
2
)
(20)
K
//
(x) =
C
0
x
q
sin(
pq
2
)[1 kx
r
cos(
pr
2
)[ C
0
kx
qr
sin(
pr
2
) cos(
pq
2
)
1 k
2
x
2r
2kx
r
cos(
pr
2
)
(21)
Values for the damper parameters, C
0
= 15; 000 Ns/m, k = 0:3 s
0:6
, r = 0:6 and q = 1, were determined by
Makris and Constantinou (1991) by tting Eqs. (20) and (21) to the experimental data. The resulting FDM
analytical ts are shown in Fig. 3(a) and (b).
4.2. K
/
(x) and K
//
(x) by the standard mechanical model
Using SMM (the generalized Maxwell model), the storage and loss stiness functions of a viscous
damper can be expressed as
8074 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
K
/
(x) =

m
i=1
x
2
q
2
i
K
i
x
2
q
2
i
1
(22)
K
//
(x) =

m
i=1
xq
i
K
i
x
2
q
2
i
1
(23)
where the constants K
i
, q
i
and m have the meanings similar to those of the parameters used in Eqs. (11) and
(12). Note that, for a viscoelastic liquid, the equilibrium stiness vanishes, i.e., K
e
= lim
t
K(t) =
lim
x0
K
/
(x) = 0 (Ferry, 1980). The coecients K
i
(i = 1; . . . ; m) can be found by tting Eq. (22) or
Fig. 3. Complex stiness for the viscous damper in the frequency domain; (a) storage stiness, (b) loss stiness.
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8075
Eq. (23) to available experimental data. A nine-term (m = 9) Prony series representation is employed in the
current example. Relaxation times with half-decade intervals, q
i
= 10
(i7)=2
(i = 1; . . . ; 9), are selected so that
the entire frequency range of the data may be covered by the analytical representation. One-decade spacings
of q
i
are usually adequate for many viscoelastic media (Schapery, 1974); however, for highly viscoelastic
media such as viscoelastic dampers, spacings closer than one decade are often required to accurately de-
scribe the data. If the spacing between q
i
s becomes too small, oscillations occur in the tted curve and
deviations from the original data increase. Since the number of observations (or data points) is greater than
the number of unknowns (m = 9 in our case), a least squares method can be employed. Although only one
set of data, either for K
/
or K
//
, is sucient to determine K
i
s, both data sets are used here to enhance the
quality of the t. The best-t K
i
s are found by minimizing the following functional representing the
normalized square errors:
E =

N
k=1
K
/
(x
k
)
K
/
k
_
_
_
1
_
2

K
//
(x
k
)
K
//
k
_
1
_
2
_
_
(24)
where K
/
k
and K
//
k
denote the experimental data at the circular frequency of x
k
, and N is the total number
of data points available. The model constants thus found are presented in Table 1, and K
/
and K
//
curves
represented by SMM are shown in Fig. 3(a) and (b). Both FDM and SMM represent the data very well.
4.3. K(t)
A signicant advantage of the use of SMM is that, once a linear viscoelastic material (or system)
function is established in one domain, the equivalent functions in other domains are automatically given.
For instance, the time-domain relaxation stiness, K(t), which relates the force history to the input dis-
placement history in a manner similar to Eq. (1), can be obtained from Eq. (3) with Es replaced by Ks and
setting K
e
= 0. The values for the parameters, K
i
and q
i
, presented in Table 1 are used here again. Fig. 4
shows the graphical representation of K(t) thus obtained. As long as the parameters K
i
and q
i
in Table 1
correctly represent the rheological properties of the viscous damper considered, the K(t) curve shown in
Fig. 4 represents the exact relaxation stiness of the damper. Recently, Schapery and Park (1999) developed
an approximate method of analytical interconversion between linear viscoelastic material functions based
on a comprehensive study of the weighting functions involved in the mathematical interrelationships be-
tween the functions. For instance, K(t) can be obtained from K
/
(x) by
Table 1
SMM constants for stiness functions of the viscoelastic liquid damper
i q
i
(s) K
i
(kN/m)
1 1:00E 03 5:01E 02
2 3:16E 03 1:66E 02
3 1:00E 02 1:12E 02
4 3:16E 02 8:12E 01
5 1:00E 01 3:51E 01
6 3:16E 01 1:02E 01
7 1:00E 00 2:97E 00
8 3:16E 00 5:84E 01
9 1:00E 01 1:23E 01
K
e
= 0 kN/m
8076 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
K(t)
1
k
K
/
(x)

x=1=t
(25)
where k = C(1 n) cos(np=2); the symbol C() denotes the Gamma function and n is the slope of the
source function on a loglog scale, i.e., n = d( log K
/
)=d( log x) at x = 1=t. No analytical expression of
K
/
(x) is required in Eq. (25) and the experimental data for K
/
(x) can directly be used to nd the K(t) curve.
The result of an approximate interconversion according to Eq. (25) is also shown in Fig. 4 and is seen to
practically coincide with the exact representation. A similar approximate interrelationship was given by
Schapery and Park (1999) for K
//
(x) K(t) conversion. Note that, while SMM readily gives K(t) once the
model is calibrated using the experimental data for K
/
(x) or K
//
(x), FDM does not provide an explicit
analytical expression for K(t) for general orders of fractional derivative. Only for a special case of q = 1 and
r = 0:5, an explicit expression for K(t) in terms of the material parameters used in Eq. (17) is available
(Makris, 1991).
4.4. Interconversion
A exibility function of the viscous damper, relating a displacement response to a force input, may now
be found from a stiness function that is already established, using an interrelationship between the two
functions. It is known that, when both the source and the target functions are expressed in Prony series, the
interconversion can be carried out very eciently (Park and Schapery, 1999). Specically, in view of Eq. (3)
and Eqs. (10)(16), if one set of constants, either q
i
; E
i
(i = 1; . . . ; m) and E
e
or s
j
; D
j
(j = 1; . . . ; n);
D
g
and g
0
, is known, the other set of unknown constants can be determined from a relationship, Eqs.
(A.2), (A.13) or Eq. (A.18). In our case, we have already determined a set of constants,
q
i
; K
i
(i = 1; . . . ; m), for the stiness functions as shown in Table 1 and we seek to nd a set of constants,
s
j
; L
j
(j = 1; . . . ; n); L
g
and g
0
, for the exibility functions. Again, here we deal with stiness (K) and
exibility (L) functions that are parallel with modulus (E) and compliance (D) functions, respectively. Note
that K
e
= 0 for viscoelastic uid dampers. Following the procedure introduced by Park and Schapery
(1999) and using the interrelationship (A.18) which, of the three alternate interrelationships, renders the
simplest system of equations, the problem of interconversion reduces to solving the following system of
linear algebraic equations for unknowns, L
j
(j = 1; . . . ; n):
[A[L = B or A
kj
L
j
= B
k
(summed on j; j = 1; . . . ; n; k = 1; . . . ; p) (26)
Fig. 4. Relaxation stiness for the viscous damper in the time domain.
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8077
where
A
kj
=

m
i=1
s
k
q
i
K
i
1 s
k
q
i
_ _
1
1 s
k
s
j
_ _
(j = 1; . . . ; n; k = 1; . . . ; p) (27)
and
B
k
= 1

m
i=1
s
k
q
i
K
i
1 s
k
q
i
_ _
1

m
i=1
K
i
_

1
s
k

m
i=1
q
i
K
i
_
(k = 1; . . . ; p): (28)
The symbol s
k
(k = 1; . . . ; p) denote the discrete values of the Laplace-transform parameter at which the
interrelationship is established. Time constants s
j
(j = 1; . . . ; n) can be either determined following a nu-
merical procedure (Park and Schapery, 1999) or specied manually from experience to avoid dealing with a
system of nonlinear equations with 2n unknowns. The number of sampling points (or the number of
equations) should not be less than the number of unknowns (i.e., p Pn). The collocation method is eected
when p = n and the least squares method can be used when p > n. In the case of the least squares method,
a minimization of the square error |B [A[L|
2
with respect to L
j
(j = 1; . . . ; n) leads to the re-
placement of [A[L = B in Eq. (26) with [A[
T
[A[L = [A[
T
B in which the product [A[
T
[A[ is a square
matrix. In our example, a set of retardation time constants s
j
(j = 1; . . . ; 8) are numerically determined
following the procedure given by Park and Schapery (1999) and is tabulated in Table 2 together with the
results for L
j
(j = 1; . . . ; 8), L
g
and g
0
. The least squares method is used to solve the system of equations. It
is to be noted that when a stiness function is represented by FDM, it cannot readily be converted into the
corresponding exibility function. For the details of the interconversion procedure used in the current and
next examples, the reader is referred to Park and Schapery (1999).
4.5. L
/
(x) and L
//
(x)
The storage and loss exibility functions for a viscous damper can be represented by SMM (the gen-
eralized Voigt model) as
L
/
(x) = L
g

n
j=1
L
j
x
2
s
2
j
1
(29)
L
//
(x) =
1
g
0
x

n
j=1
xs
j
L
j
x
2
s
2
j
1
: (30)
Table 2
SMM constants for exibility functions of the viscoelastic liquid damper
i s
i
(s) L
i
(m/kN)
1 1:78E 03 4:77E 04
2 5:13E 03 9:90E 04
3 1:70E 02 1:35E 03
4 6:46E 02 2:78E 03
5 2:40E 01 4:92E 03
6 8:13E 01 1:04E 02
7 2:82E 00 1:76E 02
8 9:33E 00 3:40E 02
L
g
= 1:10E 03 m/kN g
0
= 1:75E 01 kNs/m
8078 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
Eqs. (29) and (30) follow from Eqs. (13) and (14) with Ds replaced with Ls. The functions L
/
and L
//
are the
real and imaginary components of the complex exibility function L
+
that relates the displacement to the
force in the frequency domain so that
^ u(x) = L
+
(ix)
^
P(x) (31)
L
+
(ix) = L
/
(x) iL
//
(x): (32)
where ^ u and
^
P denote the Fourier transforms of u and P, respectively. The curves for L
/
and L
//
represented
by Eqs. (29) and (30) are shown in Fig. 5. The constants presented in Table 2 are used in the evaluation of
these functions.
4.6. L(t)
The time-domain exibility function, L(t), relating an applied force history to the resulting displacement
history in a manner similar to that of Eq. (A.1), can be obtained from Eq. (10) with Ds replaced with Ls.
The resulting L(t) is shown graphically in Fig. 6. Again, the constants shown in Table 2 are used. Also
Fig. 6. Creep exibility for the viscous damper in the time domain.
Fig. 5. Complex exibility for the viscous damper in the frequency domain.
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8079
shown in Fig. 6, for a comparison, is an approximate L(t) curve obtained simply by taking the reciprocal of
K(t) shown in Fig. 4. It can be clearly seen that the quasi-elastic relationship, K(t)L(t) = 1, renders sig-
nicant errors especially in the long-time behavior.
4.7.
~
K(s) and
~
L(s)
Finally, the operational stiness and exibility functions,
~
K(s) and
~
L(s), dened by Eqs. (15) and (16)
with Es and Ds replaced respectively with Ks and Ls, are shown in Fig. 7. Again, the model constants
presented in Tables 1 and 2 are used in generating these curves. One may easily check the reciprocal re-
lationship of
~
K and
~
L. These functions are useful in solving viscoelastic boundary value problems following
the Laplace-transform-based elastic-viscoelastic correspondence principle.
5. Numerical example 2 a viscoelastic solid damper
Viscoelastic solid dampers have been used in the control of vibration and noise in aircrafts and machines,
and more recently in the mitigation of wind or earthquake-induced vibration of structures. Fig. 8(a) shows
a viscoelastic solid damper (or simply, a viscoelastic damper) composed of viscoelastic layers bonded with
Fig. 7. Operational stiness and exibility for the viscous damper in the Laplace-transform domain.
Fig. 8. A viscoelastic solid damper; (a) geometry (Mahmoodi, 1969), (b) forcedisplacement hysteresis loops at f = 3 Hz and T = 21C
(Shen and Soong, 1995).
8080 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
three parallel steel plates (Mahmoodi, 1969). When incorporated into a structure subjected to dynamic
loading that induces a relative motion between the outer steel anges and the center plate, a portion of the
input energy is dissipated through shear deformation of the viscoelastic layers. The concept and theory for
damping devices that use constrained viscoelastic layers were originally developed by Kerwin (1959) and
Ross et al. (1959) and further expanded by Torvik (1980) and others. A viscoelastic solid damper of the type
shown in Fig. 8(a) was experimented by Shen and Soong (1995), and they used the MPL for analytical
modeling of the damper and calibrated the model using the experimental data from stress relaxation and
steady-state harmonic tests. Fig. 8(b) displays forcedisplacement hysteresis loops measured in a sinusoidal
test conducted at the frequency of 3 Hz and the temperature of 21C. The data presented by Shen and
Soong (1995) are used here to illustrate the modeling of the damper using SMM, then the performance of
MPL and SMM are discussed and compared.
5.1. G(t)
Fig. 9 shows the time-domain shear relaxation modulus, G(t), of the viscoelastic solid damper. Because
the experimental data were not available in a tabular form, data reconstructed from the MPL represen-
tation given by Shen and Soong (1995) are used here to calibrate the generalized Maxwell model expression
of G(t), Eq. (3), with Es replaced by Gs. Shen and Soong (1995) gave G
e
= 0:05861 MPa, G
g
= 1039:1
MPa, q = 1:26E 6 s, and n = 0:586 for an MPL representation of Eq. (8), with Es replaced by Gs. Note
that for a viscoelastic solid, G
e
is nonzero. An 18-term (m = 18) Prony series is used and the model con-
stants are determined in a manner similar to that described in the above example of a viscoelastic liquid
damper. The resulting numerical values for the Prony series constants are tabulated in Table 3. Except in
the glassy and rubbery plateau regions, relaxation times with half-decade intervals are used. Compared to
the earlier viscoelastic liquid damper, the current viscoelastic solid damper requires more series terms be-
cause the latter is dened over a broader range of time (or frequency).
5.2. G
/
(x) and G
//
(x)
The real and imaginary components of the complex shear modulus are presented in Fig. 10(a) and (b).
The expressions for G
/
and G
//
, corresponding to the MPL representation of G(t) as in Eq. (8), are given by
(Shen and Soong, 1995)
Fig. 9. Shear relaxation modulus for the viscoelastic damper in the time domain.
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8081
G
/
(x) = G
e
(G
g
G
e
)(xq)
n
C(1
_
n) cos
np
2
_
xq
_

(xq)
1n
1 n
cos
p
2
_
xq
_
_
(33)
G
//
(x) = (G
g
G
e
)(xq)
n
C(1
_
n) sin
np
2
_
xq
_

(xq)
1n
1 n
sin
p
2
_
xq
_
_
(34)
The SMM representations for G
/
and G
//
in the form of Eqs. (11) and (12) are readily available using the
constants presented in Table 3. Fig. 10(a) and (b) show that MPL and SMM representations for G
/
and G
//
match very well except in the high frequency region. It is found that Eqs. (33) and (34) do not properly
represent G
/
and G
//
in the high frequency region where xq P1. Additional corrective terms are required for
Eqs. (33) and (34) to accurately depict G
/
and G
//
in that region as pointed out in the Appendix of Shen and
Soong (1995). For comparison purposes, G
/
and G
//
are also obtained directly from G(t) curve using the
approximate interconversion method of Schapery and Park (1999) and are plotted in Fig. 10. The results of
the approximate interconversion match very well with the SMM representation for G
/
. The approximate
curve for G
//
obtained directly from G(t) shows some departures from the SMM representation, although
the overall trend of the curve agrees with that of the SMM curve. Such departures are due to some intrinsic
mathematical incompatibilities between the functions G(t) and G
//
(x) (Schapery and Park, 1999). Note the
decreasing G
//
with frequency in the low and high frequency regions, which is consistent with the following
general constraints on G
//
(Christensen, 1982):
lim
x0
G
//
(x) = 0 and lim
x
G
//
(x) = 0 (35)
Shen and Soong (1995) also addressed the use of a simplied power law, basically in the form of Eq. (7).
Although the simplied power law, Eq. (7), oers an excellent description of the rubber and transition
behavior, it does not adequately describe the short-time (or high-frequency), glassy behavior of most vi-
scoelastic materials.
Table 3
SMM constants for modulus functions of the viscoelastic solid damper
i q
i
(s) G
i
(MPa)
1 1:00E 07 1:33E 01
2 1:00E 06 2:86E 02
3 3:16E 06 2:91E 02
4 1:00E 05 2:12E 02
5 3:16E 05 1:12E 02
6 1:00E 04 6:16E 01
7 3:16E 04 2:98E 01
8 1:00E 03 1:61E 01
9 3:16E 03 7:83E 00
10 1:00E 02 4:15E 00
11 3:16E 02 2:03E 00
12 1:00E 01 1:11E 00
13 3:16E 01 4:91E 01
14 1:00E 00 3:26E 01
15 3:16E 00 8:25E 02
16 1:00E 01 1:26E 01
17 1:00E 02 3:73E 02
18 1:00E 03 1:18E 02
G
e
= 5:86E 02 MPa
8082 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
5.3. J(t), J
/
(x) and J
//
(x)
The corresponding shear compliance functions, J(t), J
/
(x) and J
//
(x), of the viscoelastic solid damper
may easily be determined from an available modulus function through interconversion when both the
source and target functions are represented by SMM. The method (Park and Schapery, 1999) used above
for the viscoelastic liquid damper may equally be used for the viscoelastic solid damper in the current
example. However, slight modications to Eqs. (26)(28) are required. The unknown constants, J
j
, for the
compliance functions may be found by solving the following set of equations:
[A[J = B or A
kj
J
j
= B
k
(summed on j; j = 1; . . . ; n; k = 1; . . . ; p) (36)
where
A
kj
= G
e
_

m
i=1
s
k
q
i
G
i
1 s
k
q
i
_
1
1 s
k
s
j
_ _
(j = 1; . . . ; n; k = 1; . . . ; p) (37)
and
B
k
= 1 G
e
_

m
i=1
s
k
q
i
G
i
1 s
k
q
i
__
G
e
_

m
i=1
G
i
_
(k = 1; . . . ; p): (38)
Note that nonzero G
e
is present in Eqs. (37) and (38). The model constants thus obtained are presented in
Table 4. Equation solution techniques similar to those used earlier are employed here. The glassy com-
pliance and zero shear-rate viscosity are determined from the following relations (Park and Schapery,
1999):
J
g
=
1
G
e

m
i=1
G
i
and g
0
=

g
0

i=1
q
i
G
i
: (39)
The resulting shear creep compliance, J(t), and the storage and loss compliance, J
/
(x) and J
//
(x), are
shown in Figs. 11 and 12, respectively. Their analytical forms are the same as those of Eqs. (10), (13) and
(14) with simple notational substitutions of Js for Ds. Function J(t) obtained from the quasi-elastic ap-
proximation is also presented in Fig. 11 for comparison. At t = 1 s, for example, J(t) obtained from the
quasi-elastic relationship is 74% greater than the exact value. The quasi-elastic approximation, although
Fig. 10. Complex modulus for the viscoelastic damper in the frequency domain; (a) storage modulus, (b) loss modulus.
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8083
useful for providing rough estimates, is not good enough for accurate characterization of material functions
for viscoelastic dampers.
5.4.
~
G(s) and
~
J(s)
Fig. 13 shows the variation of the operational modulus and compliance functions in the Laplace-
transform domain. It is informative to note that, in view of Figs. 913, the curve shapes for G(t), J
/
(x) and
~
J(s) are all similar to each other, and likewise, J(t), G
/
(x) and
~
G(s) share the same trends. These general
Table 4
SMM constants for compliance functions of the viscoelastic solid damper
i s
i
(s) J
i
(MPa
1
)
1 1:02E 07 5:50E 05
2 1:29E 06 1:55E 04
3 4:79E 06 4:42E 04
4 1:70E 05 8:90E 04
5 5:62E 05 1:69E 03
6 1:91E 04 3:29E 03
7 5:89E 04 5:96E 03
8 1:95E 03 1:22E 02
9 6:03E 03 2:50E 02
10 1:95E 02 3:97E 02
11 6:03E 02 1:03E 01
12 2:00E 01 1:55E 01
13 5:75E 01 3:58E 01
14 2:09E 00 7:17E 01
15 4:79E 00 1:12E 00
16 2:57E 01 5:15E 00
17 1:62E 02 6:03E 00
18 1:23E 03 3:34E 00
J
g
= 9:64E 04 MPa
1
g
0

Fig. 11. Creep compliance for the viscoelastic damper in the time domain.
8084 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
trends provide a useful rule of thumb for the engineer to predict the response of the damper subjected to
dierent modes of excitation. Similar trends exist for viscoelastic liquid dampers.
6. Further discussions
The two examples presented above demonstrate that the SMM can accurately represent the material (or
system) functions of viscoelastic dampers and that the numerical procedure involved is simple, straight-
forward and ecient compared to other models. A Prony series is highly amenable to various mathematical
operations with its exponential basis functions. This amenable nature of a Prony series renders the pro-
cedures for calibration (or tting) of a model and interconversion between material functions computa-
tionally ecient. For example, in contrast to FDM or MPL which normally requires an involved, nonlinear
curve-tting technique, SMM can be calibrated by a standard, linear curve-tting procedure as demon-
strated above. Also, while FDM and MPL oer very limited closed-form expressions for material func-
tions in dierent domains, SMM readily provides the explicit expressions for these functions. In addition,
SMM are grounded in a well-dened physical basis. Some molecular theories indeed predict the SMM
Fig. 12. Storage and loss compliance for the viscoelastic damper in the frequency domain.
Fig. 13. Operational modulus and compliance for the viscoelastic damper in the Laplace-transform domain.
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8085
representation of linear viscoelastic material functions. Specically, the Rouses theory for dilute polymer
solutions (Rouse, 1953) and the modied Rouse theory for undiluted polymers (Ferry et al., 1955) specify
the storage and loss modulus in the form of Eqs. (11) and (12). Also, it should be noted that the constants in
the generalized Maxwell and generalized Voigt models can be chosen so that the two models are mathe-
matically equivalent, and thus a viscoelastic material depicted by one model may also be depicted by the
other. Further, Biot (1954), using the thermodynamics of linear irreversible process, showed that Eqs. (3)
and (10) are the most general representations possible for E(t) and D(t) for the isothermal case, and
Schapery (1964) similarly showed the same for certain important nonisothermal cases such as thermo-
rheologically simple behavior.
An exponential basis function for a Prony series have a very narrow range of transition behavior while a
single-term in MPL has a broad range of transition behavior extending over many decades of logarithmic
time, which explains why many terms are required for a Prony series to depict a broad-band behavior. The
function, e
t
, when plotted against log t, is transient only within the approximate range of 2 < log t < 1,
and is practically constant outside of this range (Cost and Becker, 1970). Fig. 14 shows the variations of
typical, narrow-band, exponential basis functions for a Prony series and a broadband, MPL function. For
brevity but without loss of generality, simple values are assigned to the constants involved. While an MPL
function can cover a broad range with a single term, it often fails to accurately depict the entire range of
experimental data because of its limited degrees of freedom. Whereas, a multi-term Prony series can de-
scribe the same data much more accurately with expanded degrees of freedom. As pointed out above, there
is a close interrelationship between FDM and MPL. Therefore, FDM can also describe a broad-band
viscoelastic behavior with a small number of parameters. A comprehensive discussion of various basis
functions for dierent analytical representations of linear viscoelastic material behavior is given by
Tschoegl (1989).
Finally, it is to be noted that, from the theory of linear viscoelasticity, E
/
and E
//
are not independent but
are related to each other (Tschoegl, 1989),
E
/
(x) = E
e

2x
2
p
_

0
E
//
(k)
1
k(x
2
k
2
)
dk (40)
or
Fig. 14. Behavior of sample Prony series basis functions in comparison with a MPL representation.
8086 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
E
//
(x) =
2x
p
_

0
[E
/
(k) E
e
[
1
k
2
x
2
dk (41)
Eqs. (40) and (41) are known as KronigKramers relations and the integrals are to be interpreted as Cauchy
principal values. Schapery and Park (1999), based on Eqs. (40) and (41), developed the following ap-
proximate analytical interrelationships between E
/
and E
//
:
E
/
(x) E
e

1
k
+
E
//
( ^ x)

^ x=x
(42)
E
//
(x) k
+
[E
/
( ^ x) E
e
[

^ x=x
(43)
where k
+
= tan(np=2) and n = d( log E
//
)=d( log x) when E
//
is the source function, or n = d( log(E
/
E
e
))=
d( log x) when E
/
is the source function. Relations (42) and (43) may be used by the engineer to check the
adequacy of the experimental data obtained from a steady-state harmonic test.
7. Conclusions
Dierent approaches to the mathematical modeling of the rheological behavior of viscoelastic dampers
are discussed. Their theoretical basis and performance are reviewed and compared. The classical, SMM
consisting of a combination of linear springs and dashpots is shown to be more ecient than other widely
used models such as the FDM and the MPL in the mechanical characterization of viscoelastic dampers. It
is found that, despite the widespread notion of the inadequacy of spring-dashpot mechanical models
for viscoelastic dampers, the generalized Maxwell or generalized Voigt model, with their expanded degrees
of freedom, accurately describes the broad-band rheological behavior of common viscoelastic dampers.
Some reported inadequacies are attributed to the use of unduly simple mechanical models such as those
consisting of a single Maxwell or Voigt unit. In the time domain, the SMM renders a Prony series ex-
pression for the modulus or compliance function. The outstanding mathematical eciency associated with
the exponential basis functions of a Prony series greatly facilitates numerical operations that involve the
series. The methods of tting and interconversion associated with the SMM are straightforward and e-
cient, without requiring cumbersome nonlinear regression or equation-solving procedures. The material
functions in dierent domains for viscoelastic dampers can be readily determined from experimental data
(either from quasi-static or dynamic tests) or through interconversion from a function established in an-
other domain. Numerical examples on two commonly used viscoelastic dampers illustrating the detailed
procedure for tting and interconversion demonstrate the superiority of the SMM to other prevailing
models in accuracy and computational facility.
Appendix A. Interrelationships between linear viscoelastic material functions
A.1. In the time domain
The stressstrain relation (1) is used to nd the stress response of a linear viscoelastic medium to a strain
input. Conversely, the strain response to a given stress input is given by
e(t) =
_
t
0
D(t s)
dr(s)
ds
ds (A:1)
where D(t) is the creep compliance. From Eqs. (1) and (A.1), setting, e.g., r(t) = H(t) where H(t) is the
Heaviside step function with H(t) = 1 for t > 0 and H(t) = 0 for t < 0, one nds the following integral
relationship between the relaxation modulus and creep compliance:
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8087
_
t
0
E(t s)
dD(s)
ds
ds = 1 (t > 0): (A:2)
A.2. In the frequency domain
The complex material functions arise from a steady-state harmonic input. For instance, substituting
e(t) = e
A
e
ixt
and r(t) = r
A
e
ixt
respectively into Eqs. (1) and (A.1) as input, one nds
r(t) = E
+
(ix)e(t) (A:3)
e(t) = D
+
(ix)r(t) (A:4)
where E
+
and D
+
are the complex modulus and complex compliance, respectively, and have their real and
imaginary parts such that
E
+
(ix) = E
/
(x) iE
//
(x) (A:5)
D
+
(ix) = D
/
(x) iD
//
(x): (A:6)
The real and imaginary components of these complex material functions can be expressed in terms of
transient material functions (Tschoegl, 1989) as
E
/
(x) = E
e
x
_

0
[E(t) E
e
[ sinxt dt (A:7)
E
//
(x) = x
_

0
[E(t) E
e
[ cosxt dt (A:8)
D
/
(x) = D
e
x
_

0
[D(t) D
e
[ sinxt dt (A:9)
D
//
(x) = x
_

0
[D(t) D
e
[ cosxt dt (A:10)
in which E
e
and D
e
denote the long-time equilibrium modulus and compliance, respectively; i.e., E
e
=
lim
t
E(t) and D
e
= lim
t
D(t). Note that a minus sign is used in Eq. (A.6) so that D
//
will be positive.
The real parts, E
/
and D
/
, are commonly referred to as the storage modulus and compliance, and the
imaginary parts, E
//
and D
//
, the loss modulus and compliance, respectively.
It should be noted that the elastic-like Eqs. (A.3) and (A.4) apply only to steady-state harmonic motions.
However, the same form of equations apply to transient motions when Eqs. (1) and (A.1) are Fourier
transformed, i.e.,
^ r(x) = E
+
(ix)^e(x) (A:11)
^e(x) = D
+
(ix)^ r(x) (A:12)
where ^ r and ^e denote the Fourier transforms of r and e, respectively; ^ r(x) =
_

r(t)e
ixt
dt. From Eqs.
(A.3) and (A.4), or Eqs. (A.11) and (A.12), the following relationship between the complex modulus and
compliance can be seen:
8088 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
E
+
(ix)D
+
(ix) = 1: (A:13)
A.3. In the Laplace-transform domain
Taking Laplace transforms of Eqs. (1) and (A.1), one nds
r(s) =
~
E(s)e(s) (A:14)
e(s) =
~
D(s) r(s) (A:15)
where r and e denote the Laplace transforms of r and e, respectively; r(s) =
_

0
r(t)e
st
dt. The functions
~
E
and
~
D in Eqs. (A.14) and (A.15) are often referred to as the operational modulus and operational com-
pliance, respectively, and are dened by
~
E(s) = s

E(s) (A:16)
~
D(s) = s

D(s) (A:17)
where

E and

D are the Laplace transforms of E(t) and D(t), respectively. From Eqs. (A.14) and (A.15), or
by taking the Laplace transform of Eq. (A.2), one readily nds
~
E(s)
~
D(s) = 1: (A:18)
Eqs. (A.14), (A.15), and (A.18) have forms identical to those of the corresponding elastic equations,
providing the basis for the so-called elasticviscoelastic correspondence principle. The operational modulus
and compliance, Eqs. (A.16) and (A.17), are commonly involved in solving linear viscoelastic boundary
value problems via the correspondence principle.
Further, from Eqs. (A.16), (A.17) and (A.5)(A.10), the following useful relationships between the
operational and complex material functions result (Pipkin, 1972; Tschoegl, 1989):
E
+
(ix) =
~
E(s)[
six
(A:19)
D
+
(ix) =
~
D(s)[
six
(A:20)
References
Ader, C., Jendrzejczyk, J., Mangra, D., 1995. The application of viscoelastic damping materials to control the vibration of magnets in a
synchrotron radiation facility. In: Karim-Panahi, K. (Ed.), Advances in Vibration Issues, Active and Passive Vibration Mitigation,
Damping and Seismic Isolation, PVP-vol. 309. ASME, New York, pp. 18.
Aprile, A., Inaudi, J.A., Kelly, J.M., 1997. Evolutionary model of viscoelastic dampers for structural applications. J. Engng. Mech. 123
(6), 551560.
Arima, F., Miyazaki, M., Tanaka, H., Yamazaki, Y., 1988. A study on buildings with large damping using viscous damping walls.
Ninth World Conference on Earthquake Engineering, Tokyo, V, pp. 821826.
Bagley, R.L., 1989. Power law and fractional calculus model of viscoelasticity. AIAA J. 27 (10), 14121417.
Bagley, R.L., Torvik, P.J., 1983a. Fractional calculus a dierent approach to the analysis of viscoelastically damped structures.
AIAA J. 21 (5), 741748.
Bagley, R.L., Torvik, P.J., 1983b. A theoretical basis for the application of fractional calculus. J. Rheol. 27, 201210.
Baumgaertel, M., Winter, H.H., 1989. Determination of discrete relaxation and retardation time spectra from dynamic mechanical
data. Rheol. Acta 28, 511519.
Bergman, D.M., Hanson, R.D., 1993. Viscoelastic mechanical damping devices tested at real earthquake displacements. Earthquake
Spectra 9 (3), 389417.
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8089
Biot, M.A., 1954. Theory of stress-strain relations in anisotropic viscoelasticity and relaxation phenomena. J. Appl. Phys. 25, 1385
1391.
Bland, D.R., 1960. The Theory of Linear Viscoelasticity. Pergamon Press, New York.
Chang, K.C., Soong, T.T., Lai, M.L., Nielsen, E.J., 1993. Viscoelastic dampers as energy dissipation devices for seismic applications.
Earthquake Spectra 9 (3), 371387.
Chang, K.C., Soong, T.T., Oh, S.T., Lai, M.L., 1995. Seismic behavior of steel frame with added viscoelastic dampers. J. Struct.
Engng. 121 (10), 14181426.
Christensen, R.M., 1982. Theory of Viscoelasticity, second ed. Academic Press, New York.
Cost, T.L., Becker, E.B., 1970. A multidata method of approximate Laplace transform inversion. Int. J. Numer. Meth. Engng. 2, 207
219.
Dyke, S.J., Spencer, B.F., Sain, M.K., Carlson, J.D., 1996. Modeling and control of magnetorheological dampers for seismic response
reduction. Smart Mater. Struct. 5 (5), 565575.
Dyke, S.J., Spencer Jr., B.F., Sain, M.K., Carlson, J.D., 1998. Experimental study of MR dampers for seismic protection. Smart
Mater. Struct. 7 (5), 693703.
Elster, C., Honerkamp, J., 1991. Modied maximum entropy method and its application to creep data. Macromolecules 24, 310314.
Elster, C., Honerkamp, J., Weese, J., 1991. Using regularization methods for the determination of relaxation and retardation spectra of
polymeric liquids. Rheol. Acta 30, 161174.
Emri, I., Tschoegl, N.W., 1993. Generating line spectra from experimental responses. Part I: Relaxation modulus and creep
compliance. Rheol. Acta 32, 311321.
Emri, I., Tschoegl, N.W., 1994. Generating line spectra from experimental responses. Part IV: Application to experimental data.
Rheol. Acta 33, 6070.
Emri, I., Tschoegl, N.W., 1995. Determination of mechanical spectra from experimental responses. Int. J. Solids Struct. 32, 817826.
Ferry, J.D., 1980. Viscoelastic properties of polymers, third ed. Wiley, New York.
Ferry, J.D., Landel, R.F., Williams, M.L., 1955. Extensions of the Rouse theory of viscoelastic properties to undiluted linear polymers.
J. Appl. Phys. 26 (4), 359362.
Findley, W.N., Lai, J.S., Onaran, K., 1976. Creep and Relaxation of Nonlinear Viscoelastic Materials. Dover, New York.
Fung, Y.C., 1965. Foundations of Solid Mechanics. Prentice-Hall, Englewood Clis, New Jersey.
Gavin, H.P., Hanson, R.D., Filisko, F.E., 1996a. Electrorheological dampers. 1. Analysis and design. J. Appl. Mech. 63 (3), 669675.
Gavin, H.P., Hanson, R.D., Filisko, F.E., 1996b. Electrorheological dampers. 2. Testing and modeling. J. Appl. Mech. 63 (3), 676682.
Gemant, A., 1938. On fractional dierentials. Philos. Mag. 25, 540549.
GERB, 1986. Pipework Dampers A Technical Report, GERB Vibration Control, Westmont, IL.
Hanson, R.D., 1993. Supplemental damping for improved seismic performance. Earthquake Spectra 9 (3), 319334.
Harris, C.M., Crede, C.E., 1976. Shock and Vibration Handbook, second ed. McGraw-Hill, New York.
Hayes Jr., J.R., Foutch, D.A., Wood, S.L., 1999. Inuence of viscoelastic dampers on the seismic response of a lightly reinforced
concrete at slab structure. Earthquake Spectra 15 (4), 681710.
Honerkamp, J., Weese, J., 1989. Determination of the relaxation spectrum by a regularization method. Macromolecules 22, 43724377.
Hopkins, I.L., Hamming, R.W., 1957. On creep and relaxation. J. Appl. Phys. 28 (8), 906909.
Housner, G.W., Bergman, L.A., Caughey, T.K., Chassiakos, A.G., Claus, R.O., Masri, S.F., Skelton, R.E., Soong, T.T., Spencer,
B.F., Yao, J.T.P., 1997. Structural control: past, present, and future. J. Engng. Mech. 123 (9), 897971.
Humann, G.K., 1985. Full base isolation for earthquake protection by helical springs and viscodampers. Nucl. Engng. Des. 84 (2),
331338.
Jones, D.I.J., 1980. Viscoelastic materials for damping applications. Damping Applications for Vibration Control, AMD-vol. 38.
ASME, New York, pp. 2751.
Kaschta, J., Schwarzl, F.R., 1994a. Calculation of discrete retardation spectra from creep data: I. method. Rheol. Acta 33, 517529.
Kaschta, J., Schwarzl, F.R., 1994b. Calculation of discrete retardation spectra from creep data: II. Analysis of measured creep curves.
Rheol. Acta 33, 530541.
Kerwin Jr., E.M., 1959. Damping of exural waves by a constrained viscoelastic layer. J. Acous. Soc. Am. 31 (7), 952962.
Kno, W.F., Hopkins, I.L., 1972. An improved numerical interconversion for creep compliance and relaxation modulus. J. Appl.
Polym. Sci. 16, 29632972.
Koeller, R.C., 1984. Applications of fractional calculus to the theory of viscoelasticity. J. Appl. Mech. 51, 299307.
Koh, C.G., Kelly, J.M., 1990. Application of fractional derivatives to seismic analysis of base-isolated models. Earthquake Engng.
Struct. Dyn. 19, 229241.
Li, W.Q., Tsai, C.S., 1994. Seismic mitigation of structures by using viscoelastic dampers. Nucl. Engng. Des. 147 (3), 263274.
Mahmoodi, P., 1969. Structural dampers. ASCE J. Struct. Div. 95 (8), 16611672.
Mahmoodi, P., Robertson, L.E., Yontar, M., Moy, C., Feld, L., 1987. Performance of viscoelastic dampers in World Trade Center
towers. Dynamics of Structures. In: Roesset, J.M. (Ed.), Proc. Structures Congress '87, Orlando, Florida, ASCE, New York,
pp. 632644.
8090 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092
Makris, N., 1991. Theoretical and experimental investigation of viscous dampers in applications of seismic and vibration isolation.
Ph.D. Thesis, State University of New York at Baalo, Bualo, New York.
Makris, N., 1997. Rigidity-plasticity-viscosity: Can electrorheological dampers protect base-isolated structures from near-source
ground motions? Earthquake Engng. Struct. Dyn. 26 (5), 571591.
Makris, N., Constantinou, M.C., 1991. Fractional-derivative Maxwell model for viscous dampers. J. Struct. Engng. 117 (9), 2708
2724.
Makris, N., Constantinou, M.C., 1992. Spring-viscous damper systems for combined seismic and vibration isolation. Earthquake
Engng. Struct. Dyn. 21, 649664.
Makris, N., Constantinou, M.C., 1993. Models of viscoelasticity with complex-order derivatives. J. Engng. Mech. 119 (7), 14531464.
Makris, N., Constantinou, M.C., Dargush, G.F., 1993a. Analytical model of viscoelastic uid dampers. J. Struct. Engng. 119 (11),
33103325.
Makris, N., Dargush, G.F., Constantinou, M.C., 1993b. Dynamic analysis of generalized viscoelastic uids. J. Engng. Mech. 119 (8),
16631679.
Makris, N., Dargush, G.F., Constantinou, M.C., 1995. Dynamic analysis of viscoelastic-uid dampers. J. Engng. Mech. 121 (10),
11141121.
Mead, D.W., 1994. Numerical interconversion of linear viscoelastic material functions. J. Rheol. 38, 17691795.
Morgenthaler, D.R., 1987. Design and analysis of passive damped large space structures. The Role of Damping in Vibration and Noise
Control. ASCE, New York, pp. 18.
Morland, L.W., Lee, E.H., 1960. Stress analysis for linear viscoelastic materials with temperature variation. Trans. Soc. Rheol. 4, 233.
Nutting, P.G., 1921. A new generalized law of deformation. J. Franklin Institute 191, 679685.
Park, S.W., Schapery, R.A., 1999. Methods of interconversion between linear viscoelastic material functions. Part 1-A numerical
method based on Prony series. Int. J. Solids Struct. 36 (11), 16531675.
Pipkin, A.C., 1972. Lectures on Viscoelasticity Theory. Springer, Berlin.
Ramkumar, D.H.S., Caruthers, J.M., Mavridis, H., Shro, R., 1997. Computation of the linear viscoelastic relaxation spectrum from
experimental data. J. Appl. Polym. Sci. 64, 21772189.
Rogers, L., 1983. Operators and fractional derivatives for viscoelastic constitutive equations. J. Rheol. 27 (4), 351372.
Ross, D., Ungar, E.E., Kerwin Jr., E.M., 1959. Damping of plate exural vibrations by means of viscoelastic laminae. In: Ruzicka,
J.E., (Ed.), Structural Damping. ASME, New York, pp. 4987.
Rouse Jr., P.E., 1953. The theory of the linear viscoelastic properties of dilute solutions of coiling polymers. J. Chem. Phys. 21 (7),
12721280.
Samali, B., Kwok, K.C.S., 1995. Use of viscoelastic dampers in reducing wind and earthquake-induced motion of building structures.
Engng. Struct. 17 (9), 639654.
Schapery, R.A., 1961. A simple collocation method for tting viscoelastic models to experimental data. Report GALCIT SM 61-23A,
California Institute of Technology, Pasadena, CA. November.
Schapery, R.A., 1964. Application of thermodynamics to thermomechanical, fracture, and birefringent phenomena in viscoelastic
media. J. Appl. Phys. 35, 14511465.
Schapery, R.A., 1974. Viscoelastic behavior and analysis of composite materials. In: Sendeckyj, G.P. (Ed.), In Mechanics of Composite
Materials, vol. 2, Academic Press, New York, pp. 85168.
Schapery, R.A., Park, S.W., 1999. Methods of interconversion between linear viscoelastic material functions. Part 2 An approximate
analytical method. Int. J. Solids Struct. 36 (11), 16771699.
Schwahn, K.J., Delinic, K., 1988. Verication of the reduction of structural vibrations by means of viscous dampers, ASME Pressure
Vessels and Piping Conference, PVP vol. 144, June1923, Pittsburgh, PA, pp. 8795.
Schwarzl, F.R., Struik, L.C.E., 1967. Analysis of relaxation measurements. Adv. Mol. Relaxation Processes 1, 201255.
Shen, K.L., Soong, T.T., 1995. Modeling of viscoelastic dampers for structural applications. J. Engng. Mech. 121 (6), 694701.
Shukla, A.K., Datta, T.K., 1999. Optimal use of viscoelastic dampers in building frames for seismic force. ASCE J. Struct. Engng. 125
(4), 401409.
Smit, W., deVries, H., 1970. Rheological models containing fractional derivatives. Rheol. Acta 9, 525534.
Soong, T.T., Dargush, G.F., 1997. Passive Energy Dissipation Systems in Structural Engineering. Wiley, Chichester, UK.
Sunakoda, K., Sodeyama, H., Iwata, N., Fujitani, H., Soda, S., 2000. Dynamic characteristics of magneto-rheological uid damper.
Proceedings of the seventh SPIE Smart Structures and Materials Conference, March 59, Newport Beach, CA.
Torvik, P.J., 1980. The analysis and design of constrained layer damping treatments. Damping Applications for Vibration Control,
AMD-vol. 38. ASME, New York, pp. 85112.
Tsai, C.S., 1993. Innovative design of viscoelastic dampers for seismic mitigation. Nucl. Engng. Des. 139, 165182.
Tsai, C.S., 1994. Temperature eect of viscoelastic dampers during earthquakes. J. Struct. Engng. 120 (2), 394409.
Tsai, C.S., Lee, H.H., 1993a. Applications of viscoelastic dampers to high-rise buildings. J. Struct. Engng. 119 (4), 12221233.
Tsai, C.S., Lee, H.H., 1993b. Seismic mitigation of bridges by using viscoelastic dampers. Comput. Struct. 48 (4), 719727.
Tschoegl, N.W., 1989. The Phenomenological Theory of Linear Viscoelastic Behavior. Springer, Berlin.
S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092 8091
Tschoegl, N.W., Emri, I., 1992. Generating line spectra from experimental responses. Part III: interconversion between relaxation and
retardation behavior. Int. J. Polym. Mater. 18, 117127.
Tschoegl, N.W., Emri, I., 1993. Generating line spectra from experimental responses. Part II: storage and loss functions. Rheol. Acta
32, 322327.
Williams, M.L., 1964. Structural analysis of viscoelastic materials. AIAA J. 2 (5), 785808.
Xu, Y.L., Qu, W.L., Ko, J.M., 2000. Seismic response control of frame structures using magnetorheological/electrorheological
dampers. Earthquake Engng. Struct. Dyn. 29 (5), 557575.
Yang, G., Ramallo, J.C., Spencer Jr., B.F., Carlson, J.D., Sain, M.K., 2000. Dynamic performance of large-scale MR uid dampers.
Proceedings of the 14th ASCE Engineering Mechanics Conference, May 2124, Austin, TX.
Zhang, R.H., Soong, T.T., 1992. Seismic design of viscoelastic dampers for structural applications. J. Struct. Engng. 118 (5), 1375
1392.
Zhang, R.H., Soong, T.T., Mahmoodi, P., 1989. Seismic response of steel frame structures with added viscoelastic dampers.
Earthquake Engng. Struct. Dyn. 18, 389396.
Zou, X., Ou, J., 2000. Experimental study on properties of viscoelastic dampers and seismic vibration-suppression eects of viscoelastic
damped structures. Shock Vib. Dig. 32 (1), 38.
8092 S.W. Park / International Journal of Solids and Structures 38 (2001) 80658092

Anda mungkin juga menyukai