1-s2.0-s0141029611003816-main

Upload: arjun-shende

Post on 03-Apr-2018

212 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/28/2019 1-s2.0-S0141029611003816-main

    1/13

    An experimental study on a generalized Maxwell model for nonlinear

    viscoelastic dampers used in seismic isolation

    Lyan-Ywan Lu a,, Ging-Long Lin b, Ming-Hsiang Shih c

    a Department of Construction Engineering, National Kaohsiung First University of Science and Technology, 1 University Road, Yenchao, Kaohsiung 824, Taiwanb Department of Civil Engineering, National Chung Hsing University, Taichung, Taiwanc Department of Civil Engineering, National Chi-Nan University, Puli, Nantou County, Taiwan

    a r t i c l e i n f o

    Article history:

    Received 16 May 2011

    Revised 7 September 2011

    Accepted 7 September 2011

    Available online 4 November 2011

    Keywords:

    Maxwell model

    Fluid damper

    Viscoelastic damper

    Sliding isolator

    Seismic isolation

    Energy dissipation

    Shaking table test

    a b s t r a c t

    Long-stroke fluid dampers may be installed under seismic isolation systems to provide supplementary

    damping. Due to the larger vibration amplitude and velocity, highly nonlinear viscoelastic behavior

    may exist in a long-stroke fluid damper. In order to accurately simulate the hysteretic behavior of such

    a damper, this paper presents and experimentally verifies a mathematical model called the generalized

    Maxwell model (GMM). Similar to the classic Maxwell model, the GMM is composed of a stiffness and a

    viscous elements connected in series. However, nonlinearity is incorporated into both elements of the

    GMM by assuming that their resistant forces are exponential functions of the relative velocity and defor-

    mation of the damper. By adjusting the two exponential coefficients, the GMM is able to simulate the

    more complicated viscoelastic behavior of fluid dampers. The GMM is reduced to the Maxwell model

    when both exponential coefficients are set to one. To verify the GMM, both an element test with har-

    monic excitations and a shaking table test with seismic excitations were conducted for a long-stroke fluid

    damper with highly nonlinear viscoelastic behavior. The result of the element test confirms that the

    GMM model is very accurate in simulating thehysteretic property of thefluid damper under a wide range

    of excitation frequencies, while the classic Maxwell and theviscous models may only be accurate under a

    certain excitation frequency. Moreover, the shaking table test, in which the fluid damper is used to pro-vide supplementary damping for a sliding isolation system, demonstrates that the GMM is able to more

    accurately predict the amount of energy dissipation by the damper and also the peak isolator drift of the

    isolation system, especially for an earthquake with a long-period pulse.

    2011 Published by Elsevier Ltd.

    1. Introduction

    Viscoelastic (VE) dampers generally represent a wide class of

    energy dissipation devices whose forcedisplacement relationship

    has viscous or viscoelastic mechanical properties [13]. In recent

    decades, VE dampers have been widely applied to mitigate the

    effects of vibration in civil engineering structures caused by various

    excitations, including traffic load, wind load and seismic load. Forinstance, they have been used to reduce the level of deck vertical

    vibration of high-speed rail bridges [4], mitigate the wind response

    of buildings [5], control the torsional seismic response of structures

    [6], reduce structural motion due to near-fault earthquakes [7,8],

    retrofit heritage buildings or high-tech factories under seismic load

    [9,10], suppress the large seismic displacement for rocking struc-

    tures [11], protect bridges by installing dampersat expansion joints

    [12], suppress the seismic motions of two parallel structures [13],

    and suppress the excessive isolator drift in an isolation system

    induced by near-fault earthquakes [14,15]. The effectiveness of VE

    dampers has been verified by experimental means, such as shaking

    table tests [16] or mass exciters [17]. Moreover, for practical pur-

    poses, some studies regarding design issues related to VE dampers,

    such as spectra design methods [7,18], optimal design methods

    [1921], a unified theory for both VE and Viscous dampers [1],

    and a step by step design procedure [2], etc., have also been pro-

    posed by researchers.A VE damper is usually connected to a structure through braces

    (diagonal, chevron or toggled), and is activated by the relative

    motion of the structure to which it is connected. The mechanical

    property of a VE damper may behave linearly or nonlinearly,

    depending on the constituent material (i.e., what the damper is

    made of) and the fabrication technique (how the damper is made).

    Generally speaking, a VE damper can be fabricated in either a fluid

    or solid form. A fluid-type VE damper usually consists of an orifice

    piston moving in a hollow cylinder filled with highly viscous fluid,

    whilea solid-typeVE damperconsists of a solid viscoelasticmaterial

    bonded to steel plates [3]. As compared to solid VE dampers, fluid

    dampers are able to provide longer strokes and their mechanical

    0141-0296/$ - see front matter 2011 Published by Elsevier Ltd.doi:10.1016/j.engstruct.2011.09.012

    Corresponding author. Tel.: +886 7 6011000x2127.

    E-mail address: [email protected] (L.-Y. Lu).

    Engineering Structures 34 (2012) 111123

    Contents lists available at SciVerse ScienceDirect

    Engineering Structures

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / e n g s t r u c t

    http://dx.doi.org/10.1016/j.engstruct.2011.09.012mailto:[email protected]://dx.doi.org/10.1016/j.engstruct.2011.09.012http://www.sciencedirect.com/science/journal/01410296http://www.elsevier.com/locate/engstructhttp://www.elsevier.com/locate/engstructhttp://www.sciencedirect.com/science/journal/01410296http://dx.doi.org/10.1016/j.engstruct.2011.09.012mailto:[email protected]://dx.doi.org/10.1016/j.engstruct.2011.09.012
  • 7/28/2019 1-s2.0-S0141029611003816-main

    2/13

    propertiesare lesssensitiveto thevariationof ambienttemperature.

    Therefore, fluid VE dampers have wider applications, especially in

    the field of earthquake engineering, in which the vibration ampli-

    tudes of protected structures are usuallyrelatively large when com-

    pared to other engineering applications. Additionally, some

    researchers have developed viscoplastic dampers which use visco-

    elastic or elastomeric solid materials in series connected to a yield-

    ing or friction device to control the peak damper force [22,23].

    The accurate analysis of structures with VE dampers usually re-

    lies on a mathematical model that is able to precisely capture the

    mechanical properties of the installed VE dampers. In the afore-

    mentioned studies about VE dampers, four types of classic VE mod-

    els are often employed: the linear viscous, nonlinear viscous,

    Maxwell and Kelvin models. The linear viscous model assumes that

    the damper force is directly proportional to the relative velocity

    across the damper [6,10], and the only parameter in this model

    is the damping coefficient. In the nonlinear viscous model, the

    damper force is assumed to be proportional to the power of the rel-

    ative velocity [8,18,24]; therefore, this model has two parameters,

    i.e., the damping coefficient and the power of the velocity. The

    Maxwell model is composed of stiffness (spring) and viscous

    (dashpot) elements connected in series [2527], soit has two mod-

    el parameters, i.e., the stiffness and the damping coefficients. The

    Kelvin model is similar to the Maxwell model, except that the stiff-

    ness and viscous elements are placed in parallel, and thus it also

    has two model parameters [12,19,28]. The hysteretic diagram

    resulting from these four classic VE models are an eclipse, an in-

    clined eclipse or a loop symmetric about the vertical and horizon-

    tal axes of the coordinates.

    The four classic models mentioned above, which have at most

    two model parameters, are simple to use and easily applied to

    numerically simulate the dynamic responses of structures with

    VE dampers; however, they may be accurate only for dampers

    operating at a relatively small amplitude or low frequency

    [29,30]. Some studies have found experimental evidence that with

    a longer stroke, higher velocity or more extreme conditions, the

    hysteresis loop of a fluid damper may become a twisted ellipticshape that is anti-symmetric about the origin of the hysteretic

    diagram [15,31,32]. This hysteretic behavior cannot be explained

    by the simple classic models, and thus more sophisticated models

    that consider more factors are needed. Some of the more advanced

    VE models that may involve more computational complexity are

    reviewed below.

    Black and Makris [29] conducted an experimental study on the

    viscous heating effect of fluid dampers. They concluded that under

    long-stroke motions (more than two times the piston diameter),

    only a few cycles are needed to raise the internal fluid temperature

    significantly and cause a drop in the damper force. Their study also

    proposed a cooling law to estimate the temperature change in a

    fluid damper. Based on test data of full-scale fluid dampers, Wolfe

    et al. [30] suggested that the inertia effect of a moving fluid massshould be included in the model, especially for large-size dampers

    under high-frequency or high-velocity excitations. Miyamoto et al.

    [33] developed and experimentally verified a mathematical model

    for fluid dampers under limit-state conditions. Their model was

    established in the OpenSees computer programwith the gap, hook,

    dashpot, spring and undercut elements connected in parallel and in

    series. By separating the measured damper forces, Hou et al. [32]

    found that the force associated with the stiffness element in a

    Maxwell model actually may not be linear, so they proposed a sim-

    ple forcedisplacement relationship to account for this non-linear

    stiffness effect. Based on experimental data, Mazza and Vulcano

    [5] proposed a six-element generalized model, which can be con-

    sidered as an in-parallel-combination of two Maxwell models

    and one Kelvin model, for VE dampers used in the seismic andwind response controls of steel frames. Moreover, to accurately

    capture the frequency dependence behavior of fluid dampers,

    Makris et al. [25] proposed a mathematical model using fractional

    derivatives, which has the ability to capture the viscoelastic behav-

    ior using fewer model parameters [34,35].

    As shown previously, although both simple and advanced mod-

    els for VE dampers have been proposed, it may not be possible to

    accurately represent all types of such dampers with one single

    model, because the actual mechanical properties of an installed

    VE damper may depend on several factors, including the constitu-

    ent material, fabrication details (e.g., geometric parameters), and

    operating and field conditions. To this end, the current study will

    focus on a particular problem, i.e., how to accurately model a

    fluid-type VE damper used for the protection of a seismic isolation

    system. A seismic isolation system may suffer from excessive iso-

    lator drift when subjected to a near-fault earthquake which con-

    tains strong long-period components [36,37]. Nevertheless,

    recent studies have also demonstrated that this excessive isolator

    drift can be effectively suppressed by using a damper with visco-

    elastic properties, without significantly interfering with the isola-

    tion performance [14,15,38]. Moreover, since the stroke demand

    for a damper installed under a seismic isolation system may be

    ten times that of the dampers installed inside a structural frame,

    this study will focus on fluid-type VE dampers, which are capable

    of providing a longer stroke than solid-type dampers.

    The objective of the current study is to propose and verify an

    accurate mathematical model for a long-stroke fluid VE damper

    used in conjunction with a seismic isolation system. The proposed

    model, called the generalized Maxwell model (GMM), is a modifi-

    cation of the familiar Maxwell model. The major difference be-

    tween the two models is that the GMM incorporates nonlinearity

    in both the spring and the viscous element. The resistant forces

    of both elements have fractional exponential coefficients, and by

    adjusting these coefficients GMM is able to simulate the more

    complicated viscous-elastic behavior of a fluid damper. In order

    to investigate its ability to capture the hysteretic behavior of a fluid

    damper under excitations of different frequencies, the proposed

    model will be verified experimentally by both an element testand a shaking table test. In the element test, harmonic excitations

    with different frequencies will be imposed on the specimen fluid

    damper, while in the shaking table test seismic ground motions

    with strong near-fault characteristics will be imposed on a seismi-

    cally isolated structure protected by the specimen damper.

    2. Generalized Maxwell model (GMM)

    Experimental studies have demonstrated that the resistance

    forces of some viscous dampers are not only related to the velocity

    of the damper, but also to the damper deformation. Such mechan-

    ical properties may be mathematically modeled by the generalized

    Maxwell model (GMM), as shown schematically in Fig. 1. Therestoring and viscous force components of this model are simu-

    lated by the nonlinear elastic (stiffness) element and the nonlinear

    viscous element, respectively, with these two elements connected

    in series. Mechanically, this model must satisfy the following kine-

    matic conditions

    d(t) = de(t) + dv(t)

    Elastic part Viscous part

    de(t) dv(t)

    fd(t)fd(t)

    Fig. 1. A schematic of generalized Maxwell model (GMM) for a VE damper.

    112 L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123

  • 7/28/2019 1-s2.0-S0141029611003816-main

    3/13

    dt det dvt 1:a

    _dt _det _dvt 1:b

    and also the following force condition

    fdt sgndet ke j det jne sgn _dvt cv j

    _dvt jnv 2

    where the sign function sgn is defined as

    sgnx

    1 if x > 0

    0 if x 0

    1 if x < 0

    8>: 3

    In Eqs. (1) and (2), fd(t) represents the damper (axial) force; d(t)

    denotes the total damper deformation; de(t) and dv(t) denote the

    deformations of the stiffness and viscous components, respec-

    tively; _dt, _det and_dvt represent the deformation rates (the

    relative velocities) corresponding to d(t), de(t) and dv(t); ke denotes

    the stiffness value of the stiffness element; cv is the damping

    coefficient of the viscous element; ne and nv are the exponential

    coefficients of the deformation de(t) and the velocity_dvt, respec-

    tively. If ne and ne are taken to be unity, i.e., ne nv 1, Eq. (2)

    reduces to the force equation of a classic Maxwell model, in which

    both the forces of the elastic and viscous components behave

    linearly, i.e.,

    fdt ke det cv_dvt 4

    In this study, the GMM model shown in Eq. (1)-(3) will be

    employed to simulate a nonlinear fluid VE damper. As shown

    above, there are four characteristic parameters to be determined

    in the GMM, i.e., ke, cv, nv and ne.

    3. Numerical method for the generalized maxwell model

    The damper hysteretic response resulting from the GMM model

    will be simulated and compared to test data in a latter section. Thissection will first explain the numerical method used in the simula-

    tion of the GMM. Since the hysteresis loop of a damper represents

    the relation between the damper force fd(t) and the total damper

    deformation d(t), fd(t) has to be solved as a function of d(t) first.

    To this end, Eq. (2) is rewritten as

    kecv

    sgndet jdetj

    nenv sgn _dvt j

    _dvtj 5

    The last equation can be further expressed as

    kecv

    jdetj

    nenv

    1det _dvt 6

    Substituting _dvt from Eq. (1.b) in Eq. (6) yields

    _det At det _dt 7

    where A(t) is a time-variant coefficient of the following form

    At kecv

    jdetj

    nenv

    1 8

    If the damper total deformation _dt is prescribed and treated as

    an external excitation, Eq. (7) represents the dynamic equation of

    the GMM damper excited by the prescribed disturbance _dt.

    Mathematically, this equation is also a first-order ordinary differ-

    ential equation with de(t) being the variable and A(t) being a

    time-variant coefficient, and it can be solved by many numerical

    methods. Once de(t) is solved from Eq. (7), it is substituted back

    to Eq. (2) to compute the damper force fd(t). In this study, Eq. (7)will be solved numerically by using the discrete-time state-space

    technique [39]. To do so, let us first assume that A(t) is constant

    within each computational time step, while _dt varies linearly in

    each time step, i.e.,

    At Ak; for k Dt6 t< k 1 Dt 9

    _dt _dk _dk 1 _dk

    Dtt k Dt;

    for k Dt6 t< k 1 Dt 10

    where k represents the number of the time step; Dt is the time

    interval; A[k] denotes the value of the quantity A evaluated at the

    k-th time step, i.e., at t= kDt. With the assumption of Eq. (9), within

    each time step Eq. (7) is no different from a time-invariant state-

    space equation with de(t) being the state variable. The discrete-time

    solution of this state-spaceequation can be readilyobtained and ex-

    pressed as [39]

    dek 1 Adk dek B1k_dk B2k

    _dk 1 11

    where

    Adk eAkDt 12

    B1k Ak1

    Adk 1

    DtAk

    21 Adk

    B2k Ak1

    1

    DtAk

    2Adk 1 13

    After de[k + 1] is computed from Eq. (11), it is substituted back

    to Eq. (2) to obtain the damper force at the (k + 1)-th time step, i.e.,

    fdk 1 sgndek 1 ke jdek 1jne 14

    Eqs. (11) and (14) together state that the damper force fd[k] of

    the GMM damper can be evaluated step by step in time, provided

    that the time history of the total deformation rate _dk across the

    two ends of the damper is specified. Therefore, Eqs. (11) and (14)

    can be treated as the numerical solution of the GMM and will be

    employed in the later numerical simulation.Notably, since the classic Maxwell model is a special case of the

    GMM with nv ne 1, the above numerical solution can also be

    applied to simulate the Maxwell damper. In such a case, the

    time-variant term A[k] in Eqs. (8) and (9) is reduced to a constant,

    i.e.,Ak ke=cv, and consequently the termsAd[k], B1[k] and B2[k]

    in Eqs. (12) and (13) are also reduced to constants. Therefore, the

    numerical simulation of the Maxwell model is computationally

    simpler than that of the GMM.

    4. Effect of GMM parameters on damper hysteretic behavior

    As mentioned previously, the GMM model has four characteris-

    tic parameters, and this section examines how these affect the hys-

    teretic behavior by employing the numerical method from theprevious section. This investigation can benefit the design of a fluid

    damper with the GMM property, since the designer will have a

    qualitative concept about the influence of each parameter. Fig. 2

    illustrates the changes in the damper hysteresis loop when a cer-

    tain parameter of the GMM is varied and the other three parame-

    ters are fixed. The numerical values for the four parameters used in

    the simulation are listed in Table 1. In Fig. 2, the damper force is

    simulated by using the numerical method explained in the last sec-

    tion. A harmonic displacement excitation d(t) of the following form

    is considered as applied at the two ends of the damper

    dt d0 sin2p f t 15

    where d0 and f represent the displacement amplitude and the fre-

    quency of the harmonic excitation, respectively. In the simulation,the values of d0 and fare taken to be 50 mm and 1 Hz, respectively.

    L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123 113

  • 7/28/2019 1-s2.0-S0141029611003816-main

    4/13

    (1) Effect of parameter ke: Fig. 2(a) shows that the variation of the

    damper hysteresis loop with different values of the stiffness

    coefficient ke, when the other parameters are nv ne 1

    and cv 100 (N-s/mm). In this case, the hysteretic behavior

    of the GMM model is reduced to that of a classic Maxwell

    model whose hysteresis loop is an inclined ellipse. As the stiff-ness coefficient ke decreases, the inclination of the loop

    becomes more obvious and the area of the loop declines. In

    contrast, as the stiffness increases, the loop area increasesand the Maxwell model approaches a linear viscous model

    without the stiffness effect.

    (2) Effect of parameter cv: Fig. 2(b) compares the hysteresis loops

    for various values of the viscous coefficient cv, when

    nv ne 1 and ke 1000 (N/mm). The figure shows that

    the area of the hysteresis loop is enlarged as the viscous coef-

    ficient increases. At the same time, due to the increase of the

    damper force, the loop is more inclined and the stiffness effect

    becomes more obvious.

    (3) Effect of parameter ne: Fig. 2(c) illustrates the variation of the

    hysteresis loop as the stiffness exponent ne of the GMM is

    changing, while the other parameters are fixed at nv 1 (lin-

    ear viscous force), cv 100 (N-s/mm) and ke 1000 (N/mm)

    (see Table 1). It is evident that the hysteretic behavior of thedamper is insensitive to ne, when ne is greater than one. When

    Fig. 2. Parametric study of the GMM model.

    Table 1

    Numerical values used in the parametric study of the generalized Maxwell model.Parameter name Value

    Stiffness coefficient, ke 1000 (kN=mne ) or as shown in Fig. 2(a)

    Damping coefficient, cv 100 (kN=m=snv ) or as shown in Fig. 2(b)

    Stiffness exponent, ne 1.0 or as shown in Fig. 2(c)

    Damping exponent, nv 1.0 or as shown in Fig. 2(d)

    114 L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123

  • 7/28/2019 1-s2.0-S0141029611003816-main

    5/13

    ne is reduced to less than one, the hysteresis loop becomes a

    distorted ellipse with two inclined line segments at the two

    ends, and the loop area is significantly reduced.

    (4) Effect of parameter nv: Fig. 2(d) depicts the hysteresis loop of

    the GMM with the variation of the viscous exponent nv, while

    the other three parameters are fixed at ne 1 (linear restoring

    force), cv 100 (N-s/mm) and ke 1000 (N/mm) (see Table

    1). As shown in the figure, the loop is more sensitive to the

    value ofnv, when nv is less than one. The enclosed area quickly

    diminishes when nv is reduced to less than 0.5. Moreover,

    when nv is larger than unity, the hysteresis loop becomes more

    like a parallelogram with four round corners. Two inclined line

    segments which represent the stiffness effect can be clearly

    observed at the two end of the loop.

    (5) Effect of varying ne and nv: Different from Fig. 2(c) and (d), in

    which one of the exponents nv and ne remains linear,

    Fig. 2(e) illustrates the change of the hysteresis loop when

    both coefficients nv and ne are identical and varied simulta-

    neously. In this case, both the restoring and viscous forces

    may become nonlinear. The figure shows that the enclosed

    area of the hysteresis loop grows when a higher value is

    adopted for nv and ne. When both nv and ne become greater

    than one, the hysteresis loop becomes neither an inclined

    ellipse nor a parallelogram. The hysteresis loop is transformed

    into a mango-shaped loop that has slightly-crooked tips at

    both ends. This shape does not have a symmetric axis, but

    rather is anti-symmetric about the origin, and it cannot be

    described by the classic Maxwell model (nv ne 1) or other

    viscous models.

    Furthermore, Fig. 2(d) actually represents the hysteresis loops

    of a nonlinear viscoelastic model extensively implemented in some

    well-known commercial structural analysis programs, such as the

    SAP2000 software. This model, which will be called the SAP2000

    model hereafter for convenience, generally consists of a linear

    spring (ne 1) and a nonlinear viscous element (nv1) connected

    in series, and is actually a special case of the GMM model withne 1. On the other hand, Fig. 2(e) represents the hysteresis loops

    of a more general GMM model in which both the spring and vis-

    cous elements are nonlinear (nv and ne not equal to one). The com-

    parison between Fig. 2(d) and (e) demonstrates that without the

    nonlinear spring effect, the SAP2000 model with the nonlinear vis-

    cous element alone cannot achieve a similar hysteresis loop char-

    acterized by the general GMM model.

    5. Cyclic element test with harmonic excitations

    5.1. Test setup

    A long-stroke viscous fluid damper was fabricated to verify the

    GMM model. The mechanical properties of this damper will betested by using an element cyclic test and a shaking table test.

    The element test aims to investigate the nonlinear hysteretic

    property of the damper, while the shaking table test is to demon-

    strate the dynamic response of the damper to a seismic load. This

    section reports the results of the element test, while the shaking

    table test will be discussed in the next section.

    Fig. 3 illustrates the setup of the element test. The fluid damperwas fabricated by using standard manufacture techniques and was

    filled with regular automobile oil. Fig. 4 shows a photo of the dam-

    per piston head with orifices. The number of orifices, which can be

    reduced by screw bolts, determines the resulting damper coeffi-

    cient. The maximum stroke of the damper was designed to be

    250 mm, which is much larger than that of a common viscous

    damper installed inside a seismic structure. In the test, the damper

    was excited by a hydraulic actuator (see Fig. 3) with a prescribed

    harmonic displacement in the form of Eq. (15). The magnitude of

    the harmonic displacement was set to d0 = 50 mm, while the exci-

    tation frequency was increased gradually from f= 0.6 to 1.5 Hz at

    an interval of 0.1 Hz. In other words, there were a total of 10 exci-

    tation frequencies tested, covering the primary working frequency

    range of the damper. The axial resistance force and the relative dis-placement of the damper were all recorded in the test, so the

    experimental hysteresis loop can be depicted after the test.

    5.2. Identification of the parametric values for the GMM model

    Using the experimental data of the damper hysteresis loop, this

    subsection identifies the optimal parametric values that best fit the

    test data for the GMM model, and this is achieved by using an opti-

    mization searching scheme called the generalized pattern search

    (GPS) algorithm [40]. Different from the traditional optimization

    methods that usually require information about the gradient or

    derivatives of the objective function to be minimized, the GPS algo-

    rithm solves an optimization problem by directly searching a set of

    points, called a mesh, around the current point, and then lookingfor the one whose value of the objective function is lower than

    the value of the current point. Once the point with a lower value

    Fig. 3. Setup of the damper element test.

    Fig. 4. Piston head with orifices of the tested fluid damper.

    L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123 115

  • 7/28/2019 1-s2.0-S0141029611003816-main

    6/13

    is found in the mesh, the GPS replaces the current point with this

    new point at the next step of the algorithm. Since the GPS does not

    require gradient and derivatives of the objective function, it can be

    applied to solve complicated optimization problems whose objec-

    tive function is not differentiable or even continuous.

    The optimal parametric values of the GMM model are searched

    for using the GPS algorithm to minimize the following objective

    function,J, which is defined by the root mean square of the error

    between the theoretical and experimental data,

    Minimize JX10i1

    Ji X10i1

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1

    Ni

    XNik1

    fi

    d; thek f

    id; exp

    k2

    vuut 16

    where Ji denotes the part of the objective function associated with

    the i-th excitation frequency; fi

    d; thek and fi

    d; expk denote the k-th

    data point of the damper force obtained from the experiment and

    the theoretical model for the i-th excitation frequency, respectively;

    Ni represents the total number of data points in the test cycles of the

    i-th excitation frequency. Note that in the last equation i = 110,

    since there are a total of 10 excitation frequencies in the test: 0.6,

    0.7, . . ., 1.5 Hz. Eq. (16) means that if a certain set of the GMM

    parameter values leads to the minimum value of J, it will also havethe least error between the experimental and theoretical data for

    the damper force. In other words, the optimal values of the GMM

    parameters are the ones that best describe the experimental dam-

    per force. Notably, the theoretical value fd; thek in Eq. (16) is com-

    puted by the numerical solution given in Eqs. (11) and (14).

    Moreover, since only the damper relative displacement d[k] is mea-

    sured in the element test, the relative velocity terms in Eq. (11) are

    approximated by the following difference equation

    _dk 1 1

    Dtdk 1 dk 17

    where the sampling period Dt is taken to be 0.01 s in the test.

    The second row of Table 2 lists the results of searching for theoptimal GMM parameter values and the minimized function value

    J. In addition, for the latter comparative study, the aforementioned

    GPS algorithm is also employed to determine the optimal parame-

    ter values for the other three models: the Maxwell model, the lin-

    ear viscous model and the SAP2000 model mentioned previously,

    when they are used to fit the test data. The optimal parameter

    values and the minimized function value J are also summarized

    in the third, fourth and fifth rows ofTable 2, for the Maxwell, linear

    viscous, SAP2000 models, respectively. Fig. 5 compares the value of

    Jas a function of the number of the searching steps for all models.

    As shown by Table 2 and Fig. 5, the searching algorithm efficiently

    converges on J= 109 kN for the GMM model, which is the lowest

    objective function value among the four models. This implies that

    the GMM model predicts the nonlinear damper force more pre-

    cisely than the other models.

    In addition, the last row ofTable 2 and Fig. 5 show that the min-imum value of the objective functionJand the optimal parameters

    for the SAP2000 model are converged to those of the Maxwell

    model, in which nv 1. This implies that, without the nonlinear

    stiffness effect, the optimal SAP2000 model that minimizes the

    error between the experimental and analytical data is reduced to

    the optimal Maxwell model. Furthermore, since the Maxwell and

    SAP2000 models lead to the same optimal parameters for the given

    test data, only the Maxwell model was considered in the later com-

    parative study.

    5.3. Comparison of hysteresis loops predicted by various models

    Fig. 6 compares the experimental hysteresis loops of the dam-

    per with those simulated by the GMM, Maxwell and the linear-vis-cous models, for the excitation frequencies 0.6, 0.9, 1.2 and 1.5 Hz.

    The dotted lines in Fig. 6 represent the experimental loops ob-

    tained from the element test, while the solid grey lines depict

    the theoretical loops simulated by using the parameter values

    listed in Table 2 for each model. Notably, in Fig. 6 the damper rel-

    ative displacement d[k] and velocity _dk taken from the test data

    are used in simulating the theoretical loops. Therefore, the exper-

    imental and theoretical loops have exactly the same relative

    displacement and velocity, whereas the damper force of the theo-

    retical loop is computed according to each of the models. The

    experiment data in Fig. 6 show that when under harmonic excita-

    tion the hysteresis loop of the damper specimen is olive-shaped

    under a lower excitation frequency, but this is then distorted under

    a higher frequency and transformed into more like a mango shape,

    which is anti-symmetric about the origin. Fig. 6(a) demonstrates

    Table 2

    Optimal parametric values for various models obtained by GPS searching scheme.

    Model name Equation Parameter Optimal value Index value

    Generalized Maxwell fdt sgndet ke j det jne

    sgn _dvt cv j _dvt jnv

    Stiffness coef. ke 711103kN=mne J= 109 kN

    Viscous coef. cv 242 kN=m=snv

    Stiffness exponent ne 1.92

    Damping exponent nv 2.03

    Maxwell fdt ke det cv_dvt Stiffness coef. ke 2432 kN/m J= 437 kN

    Viscous coef. cv 65.3 kN-s/m

    Linear viscous fdt cv_dt Viscous coef. cv 64.7 kN-s/m J= 492 kN

    SAP2000 model fdt ke det cv_dnvv t Stiffness coef. ke 2432 kN/m J= 437 kN

    Viscous coef. cv 65.3kN=m=snv

    Damping exponent nv 1.0

    Fig. 5. Value of the objective function J vs. number of searching steps.

    116 L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123

  • 7/28/2019 1-s2.0-S0141029611003816-main

    7/13

    that the simulated results obtained by the GMM model match very

    well with the experimental loops under all the excitation frequen-

    cies. On the other hand, Fig. 6(b) and (c) show that both the

    Maxwell and linear-viscous models lead to elliptic hysteresis loops

    that cannot match well with the experimental loops for all the fre-

    quencies, even though they have less simulation error around the

    frequency 0.9 Hz. Notably, under a higher frequency, the Maxwell

    model leads to a rotated ellipse. Fig. 6(b) and (c) also illustrate that

    both the Maxwell and the linear-viscous model overestimate the

    damper force under a lower frequency, but underestimate the force

    under a higher one. Fig. 6 thus indicates that the mechanical prop-

    erties of the tested long-stroke damper can be well characterized

    by the GMM model.

    Moreover, since the area of a hysteresis loop is equivalent to the

    amount of the energy dissipated by the damper per cycle, Table 3

    compares the area of the experimental loop per cycle to those of

    Fig. 6. Comparison of experimental and theoretical hysteresis loops for the element test.

    L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123 117

  • 7/28/2019 1-s2.0-S0141029611003816-main

    8/13

    the theoretical loops predicted by the three models under each

    excitation frequency, 0.6, 0.7, . . ., 1.5 Hz. Note that in the table

    the numbers in parentheses denote the error percentages of the

    dissipated energy predicted by each damper model. The error per-

    centage is defined as

    e Athe Aexp

    Aexp

    100% 18

    where Athe and Aexp denote the areas of the theoretical and experi-

    mental hysteresis loops, respectively. The areas of the experimental

    hysteresis loops were integrated numerically by using the trapezoi-

    dal rule that can be written as

    Aexp XNkk1

    1

    2fd;expk fd;expk 1dk 1 dk 19

    where Nk denotes the number of data points per cycle. Similarly, the

    areas of the theoretical loops Athe were also calculated by Eq. (19).

    Table 3 shows that the GMM is the most accurate model in predict-

    ing the energy dissipation for the whole frequency range consid-ered, with an average error less than 1% (e 0:7%). On the other

    hand, the energy dissipation predicted by the Maxwell and linear-

    viscous models have an average error percentage around 19%, and

    are only accurate around the mean excitation frequency 1.1 Hz. This

    is consistent with the findings of Fig. 6, in which both the Maxwell

    and the linear models overestimate and underestimate the energy

    dissipation capacity of the damper, under a lower and a higher exci-

    tation frequency, respectively. Therefore, Table 3 concludes that

    with regard to the issue of energy dissipation, the GMM model is

    still the most accurate one for fluid damper testing.

    6. Shaking table test with seismic excitations

    The previous section demonstrated that the GMM model is able

    to accurately capture the hysteretic behavior of a highly-nonlinear

    fluid damper under harmonic excitation. To further investigate the

    ability of the GMM model to simulate the dynamic response of a

    fluid damper subjected to a seismic load, a shaking table test was

    conducted. In the test, the fluid damper was installed in an FPS

    (friction pendulum system) sliding isolation system to provide

    supplementary damping to the system. The setup and the results

    of the shaking table test will be discussed in this section.

    6.1. Test setup

    (1) The isolated structure: Fig. 7(a) shows the overall setup of the

    shaking table test, which was conducted in the earthquakesimulation laboratory of the National Center for Research on

    Earthquake Engineering (NCREE, Taiwan). As shown, the

    long-stroke fluid damper was placed between the shaking

    table and the bottom of the base isolated structure. The iso-

    lated structure has a height of 1.62 m and a square plane view

    whose length and width are both 2.2 m (see Fig. 7(a)). The

    fluid damper was placed along the center line of the structure

    and parallel to the excitation direction. The isolated structure,

    whose structural properties are listed in Table 4, is an assem-

    bly of three mass layers. The mass layers are filled with lead

    blocks and are connected to each other by vertical tension rods

    through the four corners of the layers. The structure has very

    high rigidity, with an identified natural frequency of 17.5 Hz,

    so it may be considered as a rigid block. The reason for using

    such a rigid structure is to focus on the response of the dam-

    per, rather than the dynamic behavior of the super-structure.

    As shown in Fig. 7, two displacement sensors (LVDT) were

    deployed to measure the damper relative displacement and

    the isolator displacement. A load-cell to measure the damper

    force was also installed at one end of the damper. Additionally,

    accelerometers were placed on the top of the rigid block andalso on the shaking table, to measure the structural and

    ground accelerations.

    (2) The FPS isolators: Four FPS-type sliding isolators were used in

    the shaking table test. Each of the four columns of the isolated

    structure was mounted on a sliding isolator (see Fig. 7), whose

    properties are also summarized in Table 4. The radius of the

    FPS sliding spherical surface was designed to be 560 mm,

    which gives an isolation period of 1.5 s. The maximum isolator

    drift is 150 mm. The slider of each isolator is composed of

    polymer material. To identify the friction coefficient of the

    sliding isolators, an isolator element test was conducted inde-

    pendently before the shaking table test [36], and the results

    show that the isolators have a stable and typical FPS hysteresis

    loop with a friction coefficient of about 0.09.

    Additionally, for later numerical simulation, Fig. 7(b) shows the

    mathematical model of the tested system. As shown, the FPS isola-

    tion system is modeled by a linear spring and a friction element

    connected in parallel. The spring is used to simulate the resorting

    force produced by the concave sliding surfaces of the FPS isolators,

    while the friction element is used to simulate the friction force ex-

    erted on the sliding surfaces. In the simulation, the FPS stiffness is

    determined by ki M g=R, where M denotes the total mass of the

    isolated structure and R represents the radius of the sliding sur-

    faces (see Table 4) The friction coefficient li is determined by the

    constituent materials of the slider and sliding surface. Notably, kiand li are the only two parameters in modeling the FPS system,

    therefore, the accuracy of their values will affect the simulated re-sponse of the isolated system.

    Table 3

    Comparison of the areas of damper hysteresis-loops per cycle in the element test.

    Excitation frequency (Hz) Experimental value (m-kN) Theoretical value (m-kN)

    GMM Maxwell Linear viscous

    0.6 0.999 0.999 (0.0%) 1.582 (58.4%) 1.586 (58.8%)

    0.7 1.435 1.440 (0.3%) 2.022 (40.9%) 2.025 (41.1%)

    0.8 1.943 1.914 (1.5%) 2.422 (25.7%) 2.426 (25.9%)

    0.9 2.402 2.391 (0.5%) 2.772 (15.4%) 2.781 (15.8%)

    1.0 2.863 2.874 (0.4%) 3.089 (7.9%) 3.106 (8.5%)

    1.1 3.346 3.351 (0.2%) 3.374 (0.8%) 3.402 (1.7%)

    1.2 3.797 3.781 (0.4%) 3.599 (5.2%) 3.647 (4.0%)

    1.3 4.288 4.314 (0.6%) 3.899 (9.2%) 3.959 (7.7%)

    1.4 4.775 4.730 (0.9%) 4.091 (14.3%) 4.173 (12.6%)

    1.5 5.239 5.152 (1.7%) 4.286 (18.2%) 4.387 (16.3%)

    Average error (0.7%) (19.6%) (19.2%)

    Note: Numbers in parentheses denote the error percentages e of the theoretical values (see Eq. (18)).

    118 L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123

  • 7/28/2019 1-s2.0-S0141029611003816-main

    9/13

    (3) Input ground accelerations: To be more representative, the fol-

    lowing three ground excitations with very different character-

    istics were used in the shaking table test: (a) the first one,

    representing a typical far-field earthquake, is the famous

    1940 El Centro earthquake (see Fig. 8(a)). (b) The second

    one, representing a typical near-fault earthquake, is the

    ground acceleration recorded by the El Centro station (Array

    No. 6) in the 1979 Imperial Valley earthquake, California.

    Fig. 8(b) depicts the waveform of this near-fault earthquake,

    and it has an obvious pulse-like waveform between 5 and

    15 s. (c) The third one is an artificially generated pulse-like

    earthquake (see Fig. 8(c)), whose acceleration time-history is

    generated by the following function:

    ugt xp vp sinxpt; 0 6 t6 Tp 20

    where ugt denotes the ground acceleration, xp is called the pulse

    frequency, Tp ( 2p=xp) is the pulse period, vp is the velocityamplitude of the pulse in velocity domain. This pulse waveform,

    which was proposed by Makris and Chang [38], is used to simulate

    the typical pulse-component that exists in a near-fault earthquake.

    Fig. 8(c) depicts the generated pulse earthquake with Tp = 2s.

    6.2. Comparison of damper hysteresis loops

    Under the El Centro Earthquake, Fig. 9 compares the damper

    hysteresis loop obtained from the shaking table test to those sim-

    ulated by the GMM, Maxwell and linear-viscous models. Similarly,

    Figs. 10 and 11 compare the hysteresis loops for the Imperial Val-

    ley and the artificial-pulse earthquakes, respectively. To compare

    the energy dissipated by the damper, Table 5 calculates the areas

    of the damper experimental and theoretical hysteresis loops in-

    duced by all three types of earthquakes. Note that in the simulation

    of Figs. 911 and Table 5, the following numerical measures are

    adopted: (1) in order to compute the nonlinear time-history re-

    sponse of the isolated system that involves the super-structure,the FPS-isolators and the fluid damper, the nonlinear numerical

    method proposed by Lu et al. [41] was employed. This method,

    developed on the basis of discrete-time state-space formulation,

    is able to simulate the dynamic response of a seismic structure

    with various nonlinear passive control devices. The GMM dis-

    crete-time solution described in Section 3, which is also developed

    based on state-space formulation, was then incorporated into this

    numerical approach. (2) For each damper model, the parametric

    values listed in Table 2 were used. (3) For the isolators and the iso-

    lated structure, the parameters listed in Table 4 were used. (4) The

    actual, measured accelerations of the shaking table were used as

    the input ground accelerations in the simulation.

    Because the modeling and measurement uncertainties in the

    shaking table test are far more complicated than in the elementtest, Figs. 9(a), 10(a) and 11(a) show that for the GMM model the

    Fig. 7. The shaking table test.

    Table 4

    Properties of isolated structure and isolators for the shaking table test.

    Item Parameter Value

    Isolated structure

    (Rigid mass block)

    Natural frequency 17.5 Hz

    Total mass M 16.38 ton

    Height 1.62 m

    Height of mass center 1.08 m

    FPS isolators Isolation period Ti 2pffiffiffiffiffiffiffiffiR=g

    p1.5 s

    Radius of sliding curvature R 0.56m

    Friction coefficient li 0.09

    Maximum isolator displacement 0.15 m

    L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123 119

  • 7/28/2019 1-s2.0-S0141029611003816-main

    10/13

    0 10 20 30 40-4

    -3

    -2

    -1

    0

    1

    2

    3

    4

    Time (s)

    Acceleration(m/s2)

    0 10 20 30 40-6

    -4

    -2

    0

    2

    4

    6

    Time (s)

    Acceleration(m/s2)

    0 2 4 6 8-3

    -2

    -1

    0

    1

    2

    3

    Time (s)

    Acceleration(m/s2)

    (a) El Centro (b) Imperial Valley (c) Artificial pulse

    Fig. 8. Ground accelerations used in the shaking table test.

    Fig. 9. Comparison of damper hysteresis loops for the El Centro Earthquake (PGA = 0.4 g).

    Fig. 10. Comparison of damper hysteresis loops for the Imperial Valley Earthquake (PGA = 0.4 g).

    Fig. 11. Comparison of damper hysteresis loops for the artificial-pulse earthquake (Pulse period = 1.75s, PGA = 0.2 g).

    120 L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123

  • 7/28/2019 1-s2.0-S0141029611003816-main

    11/13

    consistency between the experimental and theoretical loops is not

    as good as that in the element test (see Fig. 6(a)). Nevertheless,

    when comparing the three damper models, Figs. 911 show that

    the GMM model leads to a hysteresis loop whose shape is most

    consistent with the experimental one under all types of the seismic

    excitations. As for predicting the amount of energy dissipated, Ta-

    ble 5 shows that the GMM model gives the most accurate areas of

    the hysteresis loops with an average error of 7.2%, which is about

    one-third of the Maxwell model error (21.6%) and one-fourth of

    the linear-viscous model error (28.2%). The Maxwell model is thus

    slightly more accurate than the linear-viscous model. Table 5 alsoshows that the GMM model is especially accurate for a seismic load

    with a long-period pulse-like waveform, such as the Imperial Val-

    ley and artificial-pulse earthquakes. The results ofTable 5 and Figs.

    911 together demonstrate that under seismic excitations the

    GMM is the model most capable of capturing the hysteretic behav-

    ior of a nonlinear fluid damper.

    Notably, due to the test arrangement, the damper elongation in

    Figs. 911 is actually equal to the isolator displacement in the test.

    Therefore, the relatively larger deviation observed in the damper

    hysteresis loops shown in Figs. 10 and 11 for all models are primar-

    ily due to the deviation of the simulated isolator displacement in

    the time-history analysis. The deviation of the isolator displace-

    ment can be due to the modeling or parametric errors contributed

    by both the FPS isolators and the fluid damper.

    6.3. Comparison of the isolation system responses

    The previous section focused on the capability of the three

    mathematical models with regard to simulating the hysteretic

    property of the fluid damper under seismic loads. Nevertheless,

    for engineering applications, it would be also desirable to investi-

    gate the influences of using different models on the overall

    response of the isolation system. Since the responses of the most

    engineering importance for an isolation system are the maximum

    isolator displacement (isolator drift) and the maximum super-

    structural acceleration, the experimental data of these two peak

    values will be investigated in this subsection.

    For the different ground motions, Table 6 compares the exper-

    imental and theoretical peak values of the isolator displacements

    predicted by the three models. Similarly, Table 7 compares the

    experimental and theoretical peak values of the structural acceler-

    ations. For the artificial-pulse ground motion, Figs. 12 and 13

    further compare the time-history responses of the isolator dis-

    placement and the structural acceleration, respectively. Notably,

    in Tables 6 and 7, a numbers in parentheses denote the error per-centage of a certain peak response predicted by the mathematical

    model, and the error percentage is defined similar to Eq. (18). For

    the prediction of the peak isolator displacement, Table 6 and

    Fig. 12 show that the GMM model has the least average error of

    about 10% among the three models. The artificial-pulse response

    in Table 6 shows the GMM is especially accurate for the isolator

    displacement induced by a long-period pulse-like earthquake, in

    which the other two models result in a fairly large error (more

    than 30%) for the isolator displacement. On the other hand, for

    the prediction of the peak structural acceleration, the last row of

    Table 7 shows that the three models have an equal level of the per-

    centage error. This implies that the selection among the three

    models has very little influence on the peak acceleration response

    of the isolation system. Moreover, the last row of Table 7 showsthat for all three models the averages of the acceleration errors

    are about 30%, which is higher than the displacement errors shown

    in Table 6. However, Fig. 13 illustrates that the time-history re-

    sponses simulated by all the three models are very consistent with

    the experimental one, even though the peak values of the theoret-

    ical responses have an error of about 10%. The relatively larger er-

    rors in the acceleration responses of the three models listed in

    Table 7 may be due to the measurement noise of the acceleration

    signal, which is more easily contaminated by ambient electrical

    noise.

    Table 5

    Comparison of the areas of the damper hysteresis-loops for the shaking table test.

    Earthquake name (PGA) Experimentalvalue (m-kN) Theoretical value (103 J)

    GMM Maxwell Linear viscous

    El Centro(0.4 g) 0.693 0.773 (11.5%) 0.858 (23.8%) 0.901 (30.0%)

    Imperial Valley (0.4 g) 0.210 0.191 (9.0%) 0.270 (28.6%) 0.288 (37.1%)

    Artificial pulse (0.2 g) 0.343 0.347 (1.2%) 0.300 (12.5%) 0.283 (17.5%)

    Average error (7.2%) (21.6%) (28.2%)

    Note: Numbers in parentheses denote the error percentages e of the theoretical values (see Eq. (18)).

    Fig. 12. Comparison of isolator displacements for the artificial-pulse earthquake (Pulse period = 1.75s, PGA = 0.2 g).

    L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123 121

  • 7/28/2019 1-s2.0-S0141029611003816-main

    12/13

    7. Conclusions

    To provide a more accurate analytical tool for fluid dampers

    with a highly nonlinear viscoelastic property, a mathematical mod-

    el called the generalized Maxwell model (GMM) is proposed and

    experimentally verified in this paper. Developed based on the Max-

    well model, the GMM is composed of a nonlinear spring and non-

    linear viscous elements, which are connected in series. The

    resistant forces of both nonlinear elements in the GMM are propor-

    tional to the powers of the relative displacement and velocity of

    the damper. As a result, the GMM has four parameters, i.e., thestiffness coefficient, viscous coefficient, stiffness exponent and

    damping exponent. The GMM will reduce to a Maxwell model

    when both the stiffness and damping exponents are equal to one.

    By adjusting these four parameters, the GMM model is able to sim-

    ulate fluid dampers with more complicated hysteretic behavior.

    To experimentally verify the proposed GMM model, both an

    element test and a shaking table test were conducted on a long-

    stroke fluid damper that was fabricated by using the standard

    manufacturing technique. In the element test, harmonic displace-

    ment excitations with different frequencies ranging from 0.6 to

    1.5 Hz were imposed on the damper. The results show that the

    specimen damper exhibits highly nonlinear viscoelastic behavior.

    Under a lower excitation frequency, the olive-shaped hysteresis

    loop is symmetrical about the coordinates axes; however, whenthe excitation frequency increases, its hysteresis loop gradually

    transforms into a mango shape that is anti-symmetrical about

    the origin. By using a numerical searching scheme, the optimal

    parametric values that best fit the data of the element test were

    sought for the GMM model, as well as for the classic Maxwell

    model and the linear viscous model for comparison purposes.

    The comparison between the experimental and simulated data

    demonstrates that the GMM model is able to more accurately cap-

    ture the aforementioned hysteretic behavior of the fluid damper

    within the whole range of excitation frequencies, while the other

    two models are only accurate under the mean excitation fre-

    quency. Over the tested frequency range, the amount of energy dis-sipated per cycle predicted by the GMM has an average error of less

    than 1%, while the average errors for the other two models are

    about 19%.

    Moreover, in the shakingtable test, the fluid damper was placed

    under an FPS (friction pendulum system) isolated structure system

    to collect the damper response data for seismic loads. Ground mo-

    tions with far-field and near-fault characteristics were all consid-

    ered in the shaking table test, so the exerted damper seismic

    responses can be more representative. The responses of the dam-

    per and the isolated system under the same ground motions were

    then simulated by using the GMM, Maxwell and linear-viscous

    models, separately. The comparison between experimental and

    simulated responses shows that the GMM provides the most accu-

    rate prediction with regard to the amount of the energy dissipatedby the damper under an earthquake with either far-field or near-

    Fig. 13. Comparison of structural acceleration responses for the artificial-pulse earthquake (Pulse period = 1.75s, PGA = 0.2 g).

    Table 6

    Peak isolator displacements for the shaking table test.

    Earthquake name (PGA) Experimental value (m) Theoretical value (m)

    GMM Maxwell Linear viscous

    El Centro (0.4 g) 0.026 0.029 (11.5%) 0.025 (3.8%) 0.023 (11.5%)Imperial Valley (0.4 g) 0.016 0.018 (12.5%) 0.017 (6.3%) 0.014 (12.5%)

    Artificial pulse (0.2 g) 0.032 0.030 (6.3%) 0.020 (37.5%) 0.020 (37.5%)

    Average error (10.1%) (15.9%) (20.5%)

    Note: Numbers in parentheses denote the error percentages of the theoretical values.

    Table 7

    Peak acceleration responses for the shaking table test.

    Earthquake name (PGA) Experimental value (m=s2) Theoretical value (m=s2)

    GMM Maxwell Linear viscous

    El Centro(0.4 g) 2.132 2.294 (7.6%) 2.386 (11.9%) 2.350 (10.2%)

    Imperial Valley (0.4 g) 1.767 2.952 (67.1%) 3.004 (70.0%) 2.777 (57.2%)

    Artificial pulse (0.2 g) 1.670 1.872 (12.1%) 1.874 (12.2%) 1.879 (12.5%)

    Average error (28.9%) (31.3%) (26.6%)

    Note: Numbers in parentheses denote the error percentages of the theoretical value.

    122 L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123

  • 7/28/2019 1-s2.0-S0141029611003816-main

    13/13

    fault characteristics. The GMM model is also able to predict the

    peak isolator displacement more accurately, especially for an

    earthquake with a long-period pulse waveform. Nevertheless, the

    comparative results also demonstrate that the three models have

    the same accuracy in predicting the peak acceleration of the isola-

    tion system.

    Acknowledgements

    This research was sponsored in part by the National Science

    Council of ROC. (Taiwan), through Grant No. 91-2211-E-327-005.

    The authors are grateful to the National Center for Research on

    Earthquake Engineering (NCREE, Taipei) for their technical support

    on the shaking table test, and also to Ms. Shiu-Wen Tzeng for pre-

    paring the test data and plots.

    References

    [1] Fu Y, Kasai K. Comparative study of frames using viscoelastic and viscous

    dampers. J Struct Eng (ASCE) 1998;124(5):51322.

    [2] Lee KS, Fan CP, Sause R, Ricles J. Simplified design procedure for frame

    buildings with viscoelastic or elastomeric structural dampers. Earthquake EngStruct Dynam 2005;34(10):127184.

    [3] Symans MD, Charney FA, Whittaker AS, Constantinou MC, Kircher CA, Johnson

    MW, McNamara RJ. Energy dissipation systems for seismic applications:

    current practice and recent developments. J Struct Eng (ASCE)

    2008;134(1):321.

    [4] Martnez Rodrigo MD, Lavado J, Museros P. Dynamic performance of existing

    high-speed railway bridges under resonant conditions retrofitted with fluid

    viscous dampers. Eng Struct 2010;32(3):80828.

    [5] Mazza F, Vulcano A. Control of the earthquake and wind dynamic response of

    steel-framed buildings by using additional braces and/or viscoelastic dampers.

    Earthquake Eng Struct Dynam 2011;40(2):15574.

    [6] Petti L, De Iuliis M. Torsional seismic response control of asymmetric-plan

    systems by using viscous dampers. Eng Struct 2008;30(11):337788.

    [7] Hubbard DT, Mavroeidis GP. Damping coefficients for near-fault ground

    motion response spectra. Soil Dynam Earthquake Eng 2011;31(3):40117.

    [8] Xu Z, Agrawal AK, He WL, Tan P. Performance of passive energy dissipation

    systems during near-field ground motion type pulses. Eng Struct

    2007;29(2):22436.

    [9] Chrysostomou CZ, Demetriou T, Pittas M, Stassis A. Retrofit of a church withlinear viscous dampers. Struct Control Health Monit 2005;12(2):197212.

    [10] Hwang JH, Wang SJ, Huang YN, Chen JF. A seismic retrofit method by

    connecting viscous dampers for microelectronics factories. Earthquake Eng

    Struct Dynam 2007;36(11):146180.

    [11] Roh H, Reinhorn AM. Modeling and seismic response of structures with

    concrete rocking columns and viscous dampers. Eng Struct

    2010;32(8):2096107.

    [12] Feng MQ, Kim JM, Shinozuka M, Purasinghe R. Viscoelastic dampers at

    expansion joints for seismic protection of bridges. J Bridge Eng (ASCE)

    2000;5(1):6774.

    [13] Zhu HP, Ge DD, Huang X. Optimum connecting dampers to reduce the seismic

    responses of parallel structures. J Sound Vib 2011;330(9):193149.

    [14] Providakis CP. Effect of supplemental damping on LRB and FPS seismic

    isolators under near-fault ground motions. Soil Dynam Earthquake Eng

    2009;29(1):8090.

    [15] Lu LY, Shih MH, Yeh SW, Lin GL, Tzeng SW. Experimental verification of using

    viscous damper for response mitigation of isolation systems subjected to near-

    fault earthquakes. In: Proceedings of the 11th East Asian-pacific conference on

    structural engineering and construction 2008, Taipei, Taiwan, November 1921, Paper No. 254.

    [16] Chang KC, Lin YY, Chen CY. Shaking table study on displacement-based design

    for seismic retrofit of existing buildings using nonlinear viscous dampers. J

    Struct Eng (ASCE) 2008;134(4):67181.

    [17] MinKW, Kim J, Lee SH.Vibration tests of 5-storey steel frame with viscoelastic

    dampers. Eng. Struct. 2004;26(6):8319.

    [18] Pekcan G, Mander JB, Chen SS. Fundamental considerations for the design of

    non-linear viscous dampers. Earthquake Eng Struct Dynam

    1999;28(11):140525.

    [19] Gluck N, Reinhorn AM, Gluck J, Levy R. Design of supplemental dampers for

    control of structures. J Struct Eng (ASCE) 1996;122(12):13949.

    [20] Uetani K, Tsuji M, Takewaki I. Application of an optimum design method topractical building frames with viscous dampers and hysteretic dampers. Eng

    Struct 2003;25(5):57992.

    [21] Singh MP, Verma NP, Moreschi LM. Seismic analysis and design with Maxwell

    dampers. J Eng Mech (ASCE) 2003;129(3):27382.

    [22] Ibrahim YE, Marshall J, Charney FA. A visco-plastic device for seismic

    protection of structures. J Constr Steel Res 2007;63(11):151528.

    [23] Karavasilis TL, Ricles JM, Sause R, Chen C. Experimental evaluation of the

    seismic performance of steel MRFs with compressed elastomer dampers using

    large-scale real-time hybrid simulation. Eng Struct 2011;33(6):185969.

    [24] Lin WH, Chopra AK. Earthquake response of elastic SDF systems with non-

    linear fluid viscous dampers. Earthquake Eng Struct Dynam

    2002;31(9):162342.

    [25] Makris N, Constantinou MC, Dargush GF. Analytical model of viscoelastic fluid

    dampers. J Struct Eng (ASCE) 1993;119(11):331025.

    [26] Hatada T, Kobori T, Ishida M, Niwa N. Dynamic analysis of structures with

    Maxwell model. Earthquake Eng Struct Dynam 2000;29(2):15976.

    [27] Chen YT, Chai YH. Effects of brace stiffness on performance of structures with

    supplemental Maxwell model-based brace-damper systems. Earthquake Eng

    Struct Dynam 2011;40(1):7592.

    [28] SinghMP, Chang TS. Seismic analysis of structures withviscoelastic dampers. J

    Eng Mech (ASCE) 2009;135(6):57180.

    [29] Black CJ, Makris N. Viscous heating of fluid dampers under small and large

    amplitude motions: experimental studies and parametric modeling. J Eng

    Mech (ASCE) 2007;133(5):56677.

    [30] Wolfe RW, Yun HB, Masri S, Tasbihgoo F, Benzoni G. Fidelity of reduced-order

    models for large-scale nonlinear orifice viscous dampers. Struc Control Health

    Monit 2008;15(8):114363.

    [31] Contantinou MC, Syman MD. Experimental study of seismic response of

    buildings with supplemental fluid dampers. Struct Design Tall Build

    1993;2:93132.

    [32] Hou CY, Hsu DS, Chen HY. Experimental verification of the restoring-stiffness

    concept used in fluid dampers for seismic energy dissipation. Struct Des Tall

    Special Build 2005;14(1):113.

    [33] Miyamoto HK, Gilani ASJ, Wada A, Ariyaratana C. Limit states and failure

    mechanisms of viscous dampers and the implications for large earthquakes.

    Earthquake Eng Struct Dynam 2010;39(11):127997.

    [34] Lewandowski R, Chorazyczewski B. Identification of the parameters of theKelvinVoigt and the Maxwell fractional models, used to modeling of

    viscoelastic dampers. Comput Struct 2010;88(12):117.

    [35] Hwang JS, Tsai CH, Wang SJ, Huang YN. Experimental study of RC building

    structures with supplemental viscous dampers and lightly reinforced walls.

    Eng Struct 2006;28(13):181624.

    [36] Lu LY, Shih MH, Tzeng SW, Chang Chien CS. Experiment of a sliding isolated

    structure subjected to near-fault ground motion. In: Proceedings of the 7th

    Pacific conference on earthquake engineering 2003, February 1315,

    Christchurch, New Zealand.

    [37] Jangid RS. Optimum lead-rubber isolation bearings for near-fault motions. Eng

    Struct 2007;29:250313.

    [38] Makris N, Chang SP. Effect of viscous, viscoplastic and friction damping on the

    response of seismic isolated structures. Earthquake Eng Struct Dynam

    2000;29(1):85107.

    [39] Meirovitch L. Dynamics and control of structures. New York: John Wiley &

    Sons; 1990.

    [40] MATLAB. The Math Works, Inc., Natick, Massachusetts; 2000.

    [41] Lu LY, Lin GL, Lin CH. A unified analysis model for energy dissipation devices

    used in seismic structures. Comp-Aided Civil Infrastruct Eng 2009;24(1):4161.

    L.-Y. Lu et al. / Engineering Structures 34 (2012) 111123 123