You are on page 1of 20

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS

Earthquake Engng Struct. Dyn. 2010; 39:1493–1512


Published online 20 September 2010 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/eqe.1040

Active base isolation of buildings subjected to seismic excitations

Chia-Ming Chang and Billie F. Spencer Jr ∗, †, ‡


Department of Civil and Environmental Engineering, University of Illinois at Urbana-Champaign,
Urbana, IL 61801, U.S.A.

SUMMARY
Structural control technology has been widely accepted as an effective means for the protection of structures
against seismic hazards. Passive base isolation is one of the structural control techniques to enhance the
performance of structures subjected to severe earthquake excitations. Isolation bearings employed at the
base of a structure naturally increases its flexibility, but concurrently results in large base displacements.
The combination of base isolation with active control, i.e. active base isolation, creates the possibility
of achieving a balanced level of control performance in reductions of either floor accelerations or base
displacements. Many theoretical papers have been written by researchers regarding active base isolation.
A few experiments have been performed to verify these theories; however, challenges in appropriately
scaling the structural system and modeling the complex nature of control–structure interaction (CSI) have
limited the applicability of these results. This paper presents the development and experimental verification
of an active base isolation system for a seismically excited building. First, the general problem formulation
and control design procedure are provided. Subsequently, the experimental setup is described; unique
features include low-friction pendular bearings and custom-manufactured low-force hydraulic actuators. A
new system identification procedure that can effectively capture the phenomena of CSI is then presented
and used to realize control-oriented models of the system. H2 /LQG control strategies employing different
performance objectives are developed and experimentally evaluated on a six degree-of-freedom shake
table in the Smart Structures Technology Laboratory at the University of Illinois at Urbana-Champaign.
The proposed control strategies are shown to perform effectively for a wide range of seismic excitations.
Copyright q 2010 John Wiley & Sons, Ltd.

Received 17 November 2009; Revised 23 April 2010; Accepted 7 June 2010

KEY WORDS: structural control; base isolation; active control

1. INTRODUCTION

Natural hazards cause numerous deaths and cost society tens of billions of dollars each year. One
of the major hazards, earthquakes, have resulted in tremendous economic and societal devastation

∗ Correspondence to: Billie F. Spencer Jr, Department of Civil and Environmental Engineering, University of Illinois
at Urbana-Champaign, Urbana, IL 61801, U.S.A.

E-mail: bfs@illinois.edu

Nathan M. and Anne M. Newmark Endowed Chair of Civil Engineering Director.

Copyright q 2010 John Wiley & Sons, Ltd.


1494 C.-M. CHANG AND B. F. SPENCER JR

in recent years. For example, the May 12, 2008, Wenchuan, China earthquake with a magnitude
of 7.9 killed more than 69 000 people and destroyed thousands of structures. Such severe losses
drive researchers to continue working to more effectively protect our civil infrastructure, and
consequently save lives, before the unrelenting forces of nature strike again.
Passive base isolation is one solution that has proven effective for enhancing structural perfor-
mance against seismic events. The main concept behind passive base isolation is to increase the
structure’s flexibility, thus avoiding potentially dangerous seismic ground motions [1–4]. As the
cost-effective manners introduce added flexibility [5], many buildings has been installed with base
isolation bearings for seismic protection, e.g. Salt Lake City and County Building in 1989, Los
Angles City Hall in 1990, San Francisco City Hall in 1999, Passadena City Hall in 2004, etc.
However, large base displacements resulting from the increased flexibility of the passive isolation
system can potentially exceed the prescribed limit of structural designs under severe seismic exci-
tations [6–8]. The revisions to the Uniform Building Code [9] mandate the consideration of such
large base displacements, which often results in the need for additional damping devices [5, 10].
Adding extra damping devices to an isolation system is a feasible approach to mitigate excessive
base displacements [11].
Active base isolation, consisting of a passive isolation system combined with control actuators,
has been proposed as an alternative means to address the drawbacks of passive isolation systems.
The efficiency offered by the base isolation system in reducing interstory drifts and floor acceler-
ations can be combined with the adaptive nature of the active control system to provide improved
performance for a wide range of earthquakes. This attractive combination has drawn the attention
of researchers to investigate on the effectiveness of this class of systems [1]. Inaudi and Kelly [12],
Nagarajaiah et al. [13], and Yang et al. [14] also evaluated active base isolation systems exper-
imentally. These experiments successfully implemented active base isolation on structures using
the developed controllers based on the analytical models; however, these experiments employed
actuators that were not appropriately scaled for the structural model, resulting in a system that had
far too much authority. Moreover, they neglected the phenomena of control–structure interaction
(CSI), which was subsequently shown to be critical to development of high-performance control
systems [15].
Researchers have also considered semi-active base isolation systems, which consist of a passive
isolation system combined with the controllable semi-active damping devices, such as MR dampers
[5, 10, 16–20]. The semi-active devices replace the active devices in base isolation systems,
providing a number of attractive features. While these studies demonstrated the applicability of
the semi-active isolation systems, the performance was less than that of the corresponding active
isolation systems.
This paper presents the development and experimental verification of an active base isolation
system for seismically excited buildings. First, the general problem formulation and control design
procedure are provided. Subsequently, the experimental setup is described; unique features include
low-friction pendular bearings and custom-manufactured low-force hydraulic actuators. A new
system identification procedure that can effectively capture the phenomena of CSI [15] is then
presented and used to realize control-oriented models of the system. H2 /LQG control strategies [21]
employing different performance objectives are developed. To account for the frequency-dependent
nature of the earthquake excitation in the control design, a Kanai-Tajimi shaping filter is incorpo-
rated into the system model [5, 22]. The efficacy of the approach is experimentally demonstrated
in both the time and the frequency domain using a six degree-of-freedom shake table in the Smart

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
ACTIVE BASE ISOLATION OF BUILDINGS SUBJECTED TO SEISMIC EXCITATIONS 1495

Structures Technology Laboratory at the University of Illinois at Urbana-Champaign. The proposed


control strategies are shown to perform effectively for a wide range of seismic excitations.

2. PROBLEM FORMULATION

To achieve high performance, a control-orientated mathematical model for the active base isolation
system should be determined using appropriate system identification techniques. Many system
identification techniques are parametric in nature, seeking to determine or update physical quan-
tities, such as mass, damping, and stiffness. For control purposes, specific values for the physical
parameters of the structural system are not required; rather, an effective model that can accu-
rately represent the dynamic relationship between the various system inputs and the outputs is
of need.
Consider, for example, the active base isolated building shown in Figure 1. The inputs to the
system are the command to the actuator and the ground acceleration; the outputs correspond to
the measurements collected by the various sensors on the building. The controller senses the
structural responses and generates the command voltage to be sent to the control devices based on
the pre-programmed control strategy. Assuming that a linear time-invariant representation can be
employed, the continuous-time structure/control system can be written as

ẋc = Ac xc +Bc u +Ec ẍg

xn

xn−1
~
~
n stories

x2

x1

LVDT

xb
xg

Figure 1. Illustration of a multi-input and multi-output active base isolation system.

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
1496 C.-M. CHANG AND B. F. SPENCER JR

y = C y xc +D y u +F y ẍg +v (1)

z = Cz xc +Dz u

where Ac , Bc , Ec , C y , D y , F y , Cz , Dz are the continuous-time state–space matrices for the system;


xc is the continuous state vector; u is the command to the control actuator; ẍ g is ground input
accelerations; y is the sensor measurements; z is for a vector of responses used in the control
performance evaluation; and v is the measurement noise. Note that the states xc in such an input–
output model typically do not have physical meaning.
After the structure/control system is appropriately identified, the controller can be designed.
For the system in Equation (1), an H2 /LOG control algorithm [21, 23–25] is employed. Because
the system identification technique proposed in this paper depends on the discrete-time model and
because the controller is implemented digitally, Equation (1) is rewritten as
xd [k +1] = Ad xd [k]+Bd u[k]+Ed ẍg [k]

yd [k] = Cdy xd [k]+Ddy u[k]+Fdy ẍg [k]+v[k] (2)

zd [k] = Cdz xd [k]+Ddz u[k]


where Ad , Bd , Ed , Cdy , Ddy , Fdy , Cdz , Ddz are the discrete-time state–space matrices for the
system; zd is the discrete-time state vector; ydm is a vector of measured responses; ydp is a vector
of regulated responses. According to the H2 /LOG control algorithm, the control objective function
and the derived control gain are represented as
 
1 N
−1
J = lim E [(Cdz xd [k]+Ddz u[k]) Q(Cdz xd [k]+Ddz u[k])+Ru [k]]
T 2
(3)
N →∞ N k=0

u[k] = −Kd x̂d [k] (4)

where Q and R are the weightings for the regulated states and the force commands and Kd is the
optimal control gain. The estimated state vector x̂d is obtained by the Kalman filter [21] given by

x̂d [k +1] = Ad x̂d [k]+Bd u[k]+L(ydy [k]−Cdy ẑd [k]−Ddy u[k]) (5)

where L is the Kalman gain. With the adequate settings for the weighting functions in the Kalman
filter, the state for the regulator effectively estimates the structural state.
To include information about the frequency content of the earthquake in the H2 /LOG control, a
Kanai-Tajimi shaping filter [22] can be appended to the model. The details of the control designs
with a shaping filter are referred for study by Yoshioka et al. [10]. The parameters used in the
Kanai-Tajimi filter are typically assigned based on the expected dominate frequency of the ground
excitation at the building site.
By changing the weighting functions in Equation (3) and the Kalman filter, a variety of control
strategies can be generated depending on the specific control objectives. Moreover, the flexibility
and robustness of H2 /LOG control strategies allow them to readily accommodate the needs of the
application. Several different controllers will be considered in this paper.

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
ACTIVE BASE ISOLATION OF BUILDINGS SUBJECTED TO SEISMIC EXCITATIONS 1497

3. EXPERIMENTAL SETUP

An active base isolation control system in this study consists of a two-bay, two-story, steel frame
model building employing six pendular type isolation bearings and three hydraulic actuators at the
base, as shown in Figure 2. Two actuators are attached to the building base to apply forces in the
x-direction, while the third actuator can apply forces in the y-direction. Note that this study only
employs the y-actuator, whereas the other two actuators are sent zero commands.
The frame building used in this control experiment is a scale model, comprised of three 45 ×
28 ×1 steel plates for floors each weighing 360 lb. Six 100 ksi columns on each floor support the


plates. The inherent damping in the building is small. Six pendular-type bearings from WorkSafe
Technologies [26] are selected to be the passive isolation devices, providing the very low damping
and stiffness. These bearings consist of two identical conical-shaped plates with a 1” ball in-
between, as shown in Figure 3. The isolation layer is 3” in height. The bearings allow the building
to move freely in the horizontally plane. Note that these bearings can be viewed as linear devices
when the displacement at the base layer is relatively small [27].

Figure 2. Active base isolation structure/control system.

Figure 3. ISO-Base bearings.

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
1498 C.-M. CHANG AND B. F. SPENCER JR

The hydraulic actuators with the Moog G761 series servo-valves [28] are custom manufactured
by Shore Western Manufacturing [29] to meet the needs of this experiment. Previous active isolation
experiments employed actuators that were not scaled appropriately for the structural model. For
example, in Inaudi and Kelly [12], the actuator had a dynamic force rating of 8 kips, whereas
the mass was only 4000 lb. This actuator provided forces much greater than can be generated
in the corresponding full-scale application. In contrast, the actuators for this experiment were
designed with a cross-sectional area of only 1/4th of an inch, which gives them a dynamic force
rating of only 500 lbs. The total stroke for each actuator is 8.76 in. Use of these actuators allows
the laboratory-scale experiments to provide more information about the potential for full-scale
implementations.
The first three natural frequencies in the y-direction are experimentally determined to be 2.6,
10.5, 20.5 Hz without the actuators attached and 6.0, 13.8, 21.3 Hz with the actuator attached,
respectively. Note that the natural frequency of the isolated building will vary slightly with the
magnitudes of the excitations due to the nonlinear restoring force provided by the cone-shaped
isolation bearing.
A number of sensors are installed on the model building for use either in the control imple-
mentation or in the determination of the control performance. Accelerometers are located at the
edges of all floors. Each actuator has a linear variable differential transformer (LVDT) to sense the
relative displacement between the building base and the ground. The sensors for use in the control
implementation depend on the requirement of the control design, but all the sensors described
previously provide the responses for the evaluation of the control performance under different
control strategies.
Implementation of the designed control strategies for the y-actuator is performed using a
6-input-and-32-output digital controller made by the dSpace. The digital controller employs a Texas
Instruments TMS320C30 chip with 16-bit A/D and D/A converters. The high overall sampling
rates and the high computation rates in this digital controller ensure the capability for the real-time
application. The detailed features and discussion of the control implementation is provided in
Spencer et al. [25], Quast et al. [30], and Dyke et al. [31].
The experiments are conducted on a six degree-of-freedom shaking table produced by the Shore
Western Manufacturing. The table can achieve peak ground accelerations of 2 g horizontally and
1 g vertically with a 2-ton test load. The table can achieve a ±5 in stroke longitudinally, ±3 in
stroke laterally, and ±3 in vertically.

4. SYSTEM IDENTIFICATION

Prior to designing the controller for this active isolation system, a major challenge is to identify an
effective model of the structure/control system. The model needs to accurately portray the system’s
input–output behavior over the frequency range for which control is needed. According to this
definition, the model of the system can be determined either in the time domain or the frequency
domain. A model obtained by a time-domain system approach, such as the subspace system
identification [32], can describe the dynamic characteristics of systems well in the time domain,
but often cannot replicate detailed behavior in the frequency domain (e.g. zero locations). Bayard
[33] proposed a transfer function curve fitting method to identify a dynamic model. Auweraer et al.
[34] modified this method to be faster using the total least-squared (LS) algorithm. Kim et al. [35]
developed a powerful tool for the frequency-domain system identification that allowed physical

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
ACTIVE BASE ISOLATION OF BUILDINGS SUBJECTED TO SEISMIC EXCITATIONS 1499

information about zeros at the origin to be considered. While these studies provided several ways
to fit the transfer functions, the computation efficiency and the accuracy with respect to zeros
of the transfer functions are limited. To address these issues, this study introduces a new system
identification approach that identifies single-input/multi-output (SIMO) frequency domain-based
models for each input, which are subsequently combined to provide the multi-input/multi-output
(MIMO) model of the structure/control system.
Several approaches for system combination have been proposed previously. Dickinson et al. [36]
used a minimal realization method to cancel repeated modes after combining all SIMO systems
into a MIMO system. Ober [37] and Dyke et al. [24] employed a balanced realization method
to combine SIMO systems by reducing the order of systems and eliminating the noise order.
However, these methods are only adequate for systems that can be very accurately identified and
have relatively small nonlinearities; otherwise, the different natural frequencies and mode shapes
will not combine, and the identified system matrix will contain duplicate dynamics. This study
proposes a method based on the naturally invariant Markov matrix to combine all SIMO systems
into a MIMO system. The resulting system model is shown to well describe the frequency-domain
behavior over the frequency range of interest.
The details of the two steps of the proposed system identification approach are described in
detail in the remainder of this section.

4.1. SIMO system identification


The first task is to determine the SIMO model for each of the system inputs based on the measured
data. The transfer functions in the z-domain are parameterized as

Bm,l (z, h)
h m, p (z, h) = (6)
A(z, h)

where h m, p is the transfer function of a SIMO system with respect to the p-th input and the m-th
output; the roots of Bm = 0 are the zeros for the transfer function; the roots of A = 0 are the poles
of the transfer function; h denotes the coefficients of the associated polynomial functions. Note
that the order of the denominator polynomial indicates the order of the system. LS minimization
of the difference between Equation (7) and the measured transfer functions is employed, i.e.


ny 
nf
L (h) = arg min Wm (zl )(Bm, p (zl , h)− A(zl , h)h rm (zl ))2 (7)
h m=1 l=1

where ny and nf are the number of outputs and the fitting points, respectively. Wm is a weighting
function designed to provide a better fit to the poles and zeros. The superscript r denotes the
transfer functions directly obtained from the experiment. Equation (2) can be expanded into a
linear LS problem that can be solved without iterative calculation [38]. After solving the linear LS
problem based on Equation (8), some errors may exist in the identified SIMO system. Hence, the
next step for the improvement is to repeat the nonlinear LS algorithm in Equation (8) iteratively,
to further decrease the errors. As will be shown in the sequel, the resulting SIMO systems provide
an excellent match for both the poles and zeros in the identified transfer functions.
Two SIMO systems are included in this control problem: (i) the first column is derived from the
actuator command u and (ii) the second column is derived from the ground acceleration input ẍ g .

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
1500 C.-M. CHANG AND B. F. SPENCER JR

Both SIMO systems can be determined using the proposed approach. The transfer function matrix
can then be assembled based on the identified SIMO models
⎡ ⎤
h xb ,u h xb ,ẍg
⎢ ⎥
⎢h h ẍb ,ẍg ⎥
⎢ ẍb ,u ⎥
Hyd ,{u,ẍg } = [Hyd ,u Hyd ,ẍg ] = ⎢


⎥ (8)
⎢ h ẍ1 ,u h ẍ1 ,ẍg ⎥
⎣ ⎦
h ẍ2 ,u h ẍ2 ,ẍg

where yd = {xb ẍb ẍ1 ẍ2 }T , xb is the base displacement, ẍb is the base absolute acceleration, ẍ1
is the first floor absolute acceleration, and ẍ 2 is the second floor absolute acceleration. Combination
of the systems and elimination of redundant dynamics to form a MIMO system is described in the
following section.

4.2. System combination


The transfer function matrix in Equation (9) can be directly realized; however, the dynamics
contained therein will be duplicated. A method to develop a minimum realization of the system
is required. In structural control applications, the control inputs are always more important than
the disturbance (i.e. the ground acceleration in this case), because of the direct effect on the
control objective in Equation (3); the manner in which the ground acceleration enters the model
only indirectly affects control objectives. This observation is used to inform the proposed system
combination approach.
To achieve system combination, consider the invariant Markov matrix, which is defined as the
product of a linear system’s controllability and observability matrices [39]. Before employing the
system combination, each SIMO system needs to be converted into a modal canonical form using
the relationship between the transfer function and the state–space representations. Both SIMO
systems written in the state–space representation are given by

A1,1 0 B1,1
x̄d,1 [k +1] = x̄d,1 [k]+ u[k]
0 A2,1 B2,1

ȳd,1 [k] = [C1,1 C2,1 ]x̄d,1 [k]+D1 u[k]


(9)
A1,2 0 E1,2
x̄d,2 [k +1] = x̄d,2 [k]+ ẍg [k]
0 A2,2 E2,2

ȳd,2 [k] = [C1,2 C2,2 ]x̄d,2 [k]+F2 ẍg [k]

where A1,1 and A1,2 denote the controllable subsystems, and A2,1 and A2,2 denote the subsystems
of the actuator command and the ground acceleration that are uncontrollable, respectively. As an
example, in the case of updating the E1,2 matrix to form the combined system matrix, the solution

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
ACTIVE BASE ISOLATION OF BUILDINGS SUBJECTED TO SEISMIC EXCITATIONS 1501

is obtained by solving a linear algebraic problem given by:


⎡ ⎤ ⎡ ⎤
C1,2 E1,2 C1,1 Ê1,2
⎢ ⎥ ⎢ ⎥
⎢ C1,2 A1,2 E1,2 ⎥ ⎢ C1,1 A1,1 Ê1,2 ⎥
⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
2 ⎢
⎢ C1,2 A1,2 E1,2 ⎥ ⎢ C1,1 A Ê1,2 ⎥
2

⎢ ⎥=⎢ 1,1 ⎥ (10)
⎢ ⎥ ⎢ ⎥
⎢ .. ⎥ ⎢ . ⎥
⎢ . ⎥ ⎢ .. ⎥
⎣ ⎦ ⎣ ⎦
C1,2 An−1
1,2 E 1,2 C A n−1

1,1 1,1 1,2
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
A1,1 0 B1,1 Ê1,2
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
x̄d,com [k +1] = ⎣ A2,1 ⎦ x̄d,com [k]+ ⎣B2,1 ⎦ u[k]+ ⎢ ⎥
⎣ 0 ⎦ ẍg [k] (11)
0 A2,2 0 E2,2
ȳd [k] = [C1,1 C2,1 C2,2 ]x̄d,com [k]+D1 u[k]+F2 ẍg [k]

where x̄d,com is the new state vector after combination; Ê1,2 denotes the updated disturbance matrix
relative to the system matrix, which shares the same poles as both original SIMO systems; n is
the number of the shared modes. Because the first system represents the effect of control input
on the responses, it is more significant than the second system. Therefore, A1,1 is derived from
the first system; the eigenvalues to the second system are matched approximately. By using both
the SIMO system identification and the system combination, the structure/control system model
employed in this study is eventually determined for use in the control designs.

4.3. System identification results


The transfer functions for the structure/control system are obtained from each individual input based
on band-limited white noise (BLWN) tests with a 50-Hz cut-off frequency. Figure 4 demonstrates
the accuracy of the results of the combined system obtained from the two-step system identification
approach. Because the system consists of two different sensor types, such as the LVDT and the
accelerometers, the weighting functions used in Equation (8) are adjusted accordingly. While
applying the SIMO system identification in the first step, the system order for the actuator input
and the ground acceleration is 9 and 6, respectively. The odd number of the system order for the
actuator input is due to CSI; the actuator can be modeled using a single-pole transfer function.
Indeed, including the effects of CSI in the control-oriented model is critical for achieving high
performance control [15]. By applying the proposed system combination approach, the order of
the combined system consequently becomes 9. The results in Figure 4 show accurate agreement
between the experimental and the combined system transfer functions, particularly with respect to
the control input (i.e. Figure 4(a)).

5. CONTROL DESIGN PROCEDURE

After the model of the structure/control system has been successfully identified, the controller can
be designed using Equations (3)–(5). A wide variety of controllers can be considered (e.g. with
different measurements used for feedback and the incorporation of a disturbance shaping filter) to

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
1502 C.-M. CHANG AND B. F. SPENCER JR

Figure 4. Final identified model: (a) from the actuator and (b) from the ground acceleration.

meet various performance objectives. Therefore, one of the goals in this study is to demonstrate
the flexibility of control designs using the H2 /LQG control algorithm and to provide and compare
different control design approaches.
To efficiently compare the control performance among the active control cases and the zeroed
control case, a series of evaluation criteria for the control performance is provided for both the
maximum responses (J1–J5) and the RMS responses (J6–J10). J1 and J6 correspond to the base
displacement; J2 and J7 correspond to the base acceleration; J3 and J8 correspond to the first floor
acceleration; J4 and J9 correspond to the roof acceleration; and J5 and J10 correspond to the base
shear. All evaluation criteria are normalized to the zero-command case [40].
The primary objective of this active isolation system is to keep the base displacement within
prescribed limits, while simultaneously reducing the floor accelerations. After observing the system
behavior in the frequency domain, the frequency components in the range of the first natural
frequency are found to most significantly influence the base displacement. Hence, all controllers
provided in this study incorporate first-mode mitigation into the designs. The excitation shaping
filter is employed to attenuate the controller action in the higher frequency region, which can be

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
ACTIVE BASE ISOLATION OF BUILDINGS SUBJECTED TO SEISMIC EXCITATIONS 1503

Table I. Detailed description of all controllers.


Disturbance
Name of Feedback Control shaping
controller measurements objective filter
Zero-command — Always sending zeros to the —
actuators for comparison to active
control cases
Case A Accelerations of the base and the two floors Minimizing all floor accelerations No
Case B Accelerations of the base and the two floors Minimizing all accelerations Yes
Case C Accelerations of the base and the two floors Minimizing the base displacement Yes
and all accelerations

Loop Gain Input Isolated


Actuator Responses
Building

Loop Gain Loop Gain Output


Transfer Controller
Function

Figure 5. Diagram of loop-gain transfer function.

deleterious to achieving the control objective. Given these considerations, the designed controllers
can achieve a balance in control performance (i.e. reductions on the base displacement versus
reductions in the floor accelerations).
Three candidate controllers based on the H2 /LQG control algorithm are designed following
the discussions in the previous paragraph. The controllers considered are different in terms of
control objectives, whether or not a disturbance shaping filter was employed, as shown in Table I.
Note that the zero-command controller always sends zero voltage to the actuator. Once these
controllers are designed, the sampling rate for the computation of the actuator command and for
the implementation of the control loop in the digital controller is appropriately set [30].
As a check of the stability of the structure/control system, a loop-gain transfer function is
measured before implementing the controller. The loop-gain transfer function is defined by the
relationship from the actuator command input to the controller command output [24], as shown
in Figure 5. According to the definition, the loop gain should be less than one at the frequencies
where the model poorly represents the structure/control system (i.e. above 25 Hz as shown in
Figure 4). Moreover, the magnitude of the loop-gain transfer function at higher frequencies should
steadily roll off and be less than one. Therefore, a control design is considered to be acceptable
if the magnitude of the loop gain at high frequencies is less than −10 dB at frequencies greater
than 25 Hz.

6. EXPERIMENTAL RESULTS

Two different types of shake table tests are considered to verify the control designs. First, a 0–50 Hz
BLWN excitation is employed to assess the frequency domain behavior of the three control designs

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
1504 C.-M. CHANG AND B. F. SPENCER JR

and the zeroed control. Note that the zeroed control indicates the case where a zero command is
sent to the actuators. For the second type of test, two earthquake records are chosen to assess the
controller’s ability to reduce the root-mean-square (RMS) and the maximum responses in the time
domain. The step-by-step procedures employed to analyze the control performance for these tests
are provided in this section.

6.1. Characteristics of isolation bearings


First, the passive base isolation system (i.e. without the actuators attached) is examined to under-
stand the behavior of the isolation bearings. BLWN shaking table tests were employed. The natural
frequencies and damping of the passive isolation system are determined and presented in Figure 6
under different levels of excitations to investigate system nonlinearity. Note that the amplitude
index in this figure corresponds to the level of the input voltage to the shaking table, which are
proportional to the peak ground accelerations. Because the pendular bearings are conical in shape,
the associated restoring force is constant, resulting in an equivalent stiffness that decreases with
an increase in the excitation level. The results also show a significant amplitude-dependence in
the isolator damping. Control designs must be robust to such system nonlinearities.
The transfer function for the passive base isolation is compared to the zero-command case in
Figure 7. As expected, the passive control system exhibits much lower natural frequency responses
than the zero-command case. Therefore, the base displacement of the passive system can be
amplified more easily.

Figure 6. Relationship of natural frequency and damping to excitation level in passive base isolation.

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
ACTIVE BASE ISOLATION OF BUILDINGS SUBJECTED TO SEISMIC EXCITATIONS 1505

Figure 7. Floor accelerations transfer functions for the passive base isolation and zero-command cases.

Figure 8. Loop-gain transfer function of active control Case B.

6.2. Loop-gain analysis


As discussed previously, the loop-gain transfer functions should be tested before final implementa-
tion of the close-loop control. Figure 8 illustrates the loop-gain transfer function using the control
Case B. The magnitudes of the experimental and simulated loop gains match well in the regions
where the loop gain is greater than 0 dB. Because the input and output of the loop-gain transfer

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
1506 C.-M. CHANG AND B. F. SPENCER JR

Table II. Comparison on maximum and RMS responses of all control cases to
zero-command case under BLWN excitations.
Active Base displacement Base acceleration First floor Second floor
control case (in/in) (g/g) acceleration (g/g) acceleration (g/g)
Percent reduction in the transfer function magnitude at the first natural frequency with respect to the
zero-command case
Case A 97.18% 95.79% 95.61% 95.03%
Case B 94.29% 96.70% 98.66% 98.31%
Case C 98.76% 88.60% 73.76% 64.76%

Figure 9. Structural transfer functions of zero-command case and control Case B.

share the same physical unit, a magnitude above one indicates that higher control authority will be
employed at these frequencies. The loop gains for the other control cases exhibited similar behavior.

6.3. Frequency-domain analysis


The experimental control results for the base displacement and all acceleration using control
Cases A–C are provided in Table II. The results are normalized with respect to the zeroed command
case and represent the magnitude of transfer functions evaluated at the first natural frequency.
The reductions of the peak values in the transfer function at the first natural frequency should
reduce peak time-domain responses when the earthquakes contain large low-frequency components
around the first natural frequency. Therefore, the control performance under earthquake excitations
can be anticipated from frequency-domain results. For control Case A, the results show the
averaged control performance among three control cases. Control Case B achieves the best control
performance in all accelerations at the base (97% reduction in maximum value) and two floors
(98–99% reductions in maximum values). Control Case C performs the best reduction in the base
displacement (99% reduction in maximum value).
In addition, Figure 9 shows the control results of the base displacement and all accelerations
using the active control Case B, for which the control objective is to minimize all accelerations.
Control Case B will be discussed in more detail in the following paragraphs because this control

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
ACTIVE BASE ISOLATION OF BUILDINGS SUBJECTED TO SEISMIC EXCITATIONS 1507

case is representative of the overall performance in terms of reducing accelerations. 97–98%


reductions at peak of the transfer function at the first natural frequency are found in the accelerations
as compared to the zero-command case, indicating that the objective of Case B to minimize all
accelerations has been achieved. Figure 9 also demonstrates the quality of the identified model
and its predictive capability; the experimental results and the simulated results match particularly
well for the first mode. However, the magnitude of the second mode is lower in the experimental
results, due to the isolator’s nonlinear damping as shown in Figure 6.
Note that while this control design focuses on the mitigation of the floor acceleration, the
base displacement is also decreased. Hence, active control Case B is an exemplar of the control
performance for active base isolation.
Damping ratios for the first mode of the control system are also determined for each control
case by curve fitting the transfer functions. The damping ratios for the first mode are 10.5, 10.4,
20.2%, respectively, for control Cases A–C. In comparison to the zero-command case (2.0% of
damping ratio), the damping ratios of all the active control cases are increased.

6.4. Time-domain analysis


Two ground motions are considered as the excitations for the evaluation of the control performance
in the time domain. The first is the North–South component of the 1940 El Centro earthquake
record at Imperial Valley, California; the other is the North–South component of the 1995 Kobe
earthquake record at Hyogo-ken, Nanbu, Japan. Both time histories are scaled appropriately for
this structure/control system.
All the values of the J -indices are normalized with respect to the control performance of the
zero-command case and listed in Table III. For the El Centro earthquake, control Case A results
in a 24% reduction in the base acceleration. The control Case B mitigates the accelerations at two
floors (48–51% reductions in maximum responses) and (41–45% reductions in RMS responses)
and the base shear (41% reduction in maximum response and 42% reduction in RMS response)
significantly. Control Case C still exhibits significant reductions in the maximum (66% reduction)
and RMS (52% reduction) responses of the base displacement, while the floor accelerations in
both values are increased. In the event of the Kobe earthquake, similar control performance among
all the control cases is found.
The results in the event of the El Centro earthquake also show the trade-off between acceleration
levels and base displacements (see control Cases A and B in Figure 10). Both controllers focus
on the mitigation of the floor accelerations; however, the level of response reduction in the base
displacements can be traded-off against the response reduction in the floor accelerations. Control
Case A achieves a 10% RMS improvement in the base shear and the roof acceleration over Case B,
whereas the Case B produces larger base displacement (40% more) than the Case A. The basic rule
of active base isolation is to enforce the base displacement to be within the limits, while ensuring
that the base shear and the floor accelerations are minimized. For the zeroed control case, the base
displacements are extremely small; allowing a somewhat larger base displacement in exchange
for smaller acceleration levels is reasonable. Therefore, even though the control Case A achieves
better performance than Case B for the displacement, control Case B is still considered to be the
best control design for this event. Similar control performance for these two control cases is also
found for the Kobe earthquake, as shown in Figure 11.
The quality of the identified model is also assessed in the time domain. Figure 12 shows the
acceleration responses of both the experimental data and the simulation results using the active

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
1508 C.-M. CHANG AND B. F. SPENCER JR

Table III. Maximum and RMS seismic responses of control cases normalized to zero-command case.
Base Base First floor Second floor
Control displacement acceleration acceleration acceleration Base shear
case J1 J2 J3 J4 J5
Maximum responses of control case normalized to zero-command case in El Centro earthquake
A 0.752(24.8%)∗ 0.619(38.1%) 0.666(33.4%) 0.645(35.5%) 0.712(28.8%)
(0.011 in/0.015 in)† (0.68 g/0.109 g) (0.108 g/0.163 g) (0.136 g/0.211 g) (323.141l b
/453.798l b)
B 1.175(−17.5%) 0.673(32.7%) 0.495(50.5%) 0.523(47.7%) 0.593(40.7%)
(0.017 in/0.015 in) (0.073 g/0.109 g) (0.081 g/0.163 g) (0.110 g/0.211 g) (269.203 lb
/453.789 lb)
C 0.342(65.8%) 0.616(38.4%) 0.899(10.1%) 0.985(1.5%) 0.871(12.9%)
(0.005 in/0.015 in) (0.067 g/0.109 g) (0.146 g/0.163 g) (0.208 g/0.211 g) (395.354 lb
/453.789 lb)
Maximum responses of control case normalized to zero-command case in Kobe earthquake
A 0.640(36.0%) 0.799(20.1%) 0.469(53.1%) 0.439(56.1%) 0.468(53.2%)
(0.018 in/0.028 in) (0.120 g/0.151 g) (0.132 g/0.281 g) (0.176 g/0.401 g) (389.886 lb
/832.531 lb)
B 0.938(6.2%) 0.798(20.2%) 0.401(59.9%) 0.335(66.5%) 0.429(57.1%)
(0.026 in/0.028 in) (0.120 g/0.151 g) (0.113 g/0.281 g) (0.134 g/0.401 g) (357.387 lb
/832.531 lb)
C 0.446(55.4%) 0.760(24.0%) 0.608(39.2%) 0.567(43.3%) 0.565(43.5%)
(0.012 in/0.028 in) (0.115 g/0.151 g) (0.171 g/0.281 g) (0.227 g/0.401 g) (470.081 lb
/832.531 lb)
J6 J7 J8 J9 J10
RMS responses of control case normalized to zero-command case in El Centro earthquake
A 0.784(21.6%) 0.759(24.1%) 0.703(29.7%) 0.726(27.4%) 0.713(28.7%)
(3.919 e−5 in (2.600 e−4 g (4.079 e−4 g (6.224 e−4 g (1.304 lb/1.830 lb)
/4.996 e−5 in) /3.424 e−4 g) /5.798 e−4 g) /8.570 e−4 g)
B 1.268(−26.8%) 0.765(23.5%) 0.593(40.7%) 0.554(44.6%) 0.578(42.2%)
(6.334 e−5 in (2.619 e−4 g (3.440 e−4 g (4.743 e−4 g /1.830 lb)/1.830 lb)
/4.996 e−5 in) /3.424 e−4 g) /5.798 e−4 g) /8.570 e−4 g)
C 0.477(52.3%) 0.838(16.2%) 1.076(−7.6%) 1.215(−21.5%) 1.093(9.3%)
(2.385 e−5 in (2.870 e−4 g (6.238 e−4 g (1.041 e−3 g (2.000 lb/1.830 lb)
/4.996 e−5 in) /3.424 e−4 g) /5.798 e−4 g) /8.570 e−4 g)
RMS responses of control case normalized to zero-command case in Kobe earthquake
A 0.508(49.2%) 0.551(44.9%) 0.387(61.3%) 0.348(65.2%) 0.385(61.5%)
(4.614 e−5 in (2.558 e−4 g (3.491 e−4 g (4.750 e−4 g (1.105 lb/2.873 lb)
/9.068 e−5 in) /4.620 e−4 g) /9.003 e−4 g) /1.365 e−3 g)
B 0.771(22.9%) 0.573(42.7%) 0.363(63.7%) 0.304(69.6%) 0.352(64.8%)
(6.997 e−5 in (2.653 e−4 g (3.272 e−4 g (4.159 e−4 g (1.013 lb/2.873 lb)
/9.068 e−5 in) /4.620 e−4 g) /9.003 e−4 g) /1.365 e−3 g)
C 0.328(67.2%) 0.546(45.4%) 0.426(57.4%) 0.407(59.3%) 0.424(57.6%)
(2.989 e−5 in (2.535 e−4 g (3.847 e−4 g (5.563 e−4 g (1.217 lb/2.873 lb)
/9.068 e−5 in) /4.620 e−4 g) /9.003 e−4 g) /1.365 e−3 g)
∗ The numbers in parentheses indicate the percent reduction with respect to the zero-command case.
† The numbers in parentheses present the responses of the control cases and the responses of the zero-command
case, respectively.

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
ACTIVE BASE ISOLATION OF BUILDINGS SUBJECTED TO SEISMIC EXCITATIONS 1509

Figure 10. Comparison on the structural responses between (a) the control Case A and (b) the control
Case B in El Centro earthquake event.

Figure 11. Structural acceleration responses for the Kobe earthquake: (a) control
Case A and (b) control Case B.

control Case B for the Kobe earthquake. This control case achieves the best reductions in the
maximum accelerations for the top two floors and the maximum base shear. The well-matched
responses verify that the identified model used in the control design can predict the structural
responses in simulation.

7. CONCLUSIONS

This paper developed and experimentally verified an active base isolation system for a seismi-
cally excited building. First, the general problem formulation and control design procedure were

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
1510 C.-M. CHANG AND B. F. SPENCER JR

Figure 12. Comparison on the acceleration responses between the experimental data and the simulation
results using the control Case B.

described. Subsequently, the experimental setup was presented, and the unique features of the
experiment identified, including the use of low-friction pendular bearings and custom-manufactured
low-force hydraulic actuators. A new system identification procedure was proposed and shown to
effectively realize control-oriented models of the system and to capture the phenomena of CSI.
H2 /LQG control strategies employing different performance objectives were developed and exper-
imentally evaluated on a six degree-of-freedom shake table in the Smart Structures Technology
Laboratory at the University of Illinois at Urbana-Champaign. The proposed control strategies
were shown to perform effectively for a wide range of seismic excitations and to be robust against
system nonlinearities. This study demonstrates the efficacy of active base isolation in reductions
of the floor accelerations and base shear, while limiting base displacement.

ACKNOWLEDGEMENTS
The authors acknowledge National Science Foundation’s support under grants CMS 03-01140 and CMS
06-00433, Dr S. C. Liu, Program Manager, and the funding from National Science Council in Taiwan
under Grant No. NSC-095-SAF-I-564-036-TMS. The authors also thank WorkSafe Technologies for the
generous donations of several ISO-Base units. The authors acknowledge T. A. Mullenix and K. S. Park
for sharing their valuable experience.

REFERENCES
1. Kelly JM, Leitmann G, Soldatos AG. Robust control of base-isolated structures under earthquake excitation.
Journal of Optimization Theory and Applications 1987; 53(2):159–180.
2. Skinner RI, Robinson WH, McVerry GH. An Introduction to Seismic Isolation. Wiley: Chichester, England, 1993.
3. Kelly JM. Earthquake Resistant Design with Rubber (2nd edn). Springer: New York, 1997.
4. Naeim F, Kelly JM. Design of Seismic Isolated Structures: From Theory to Practice. Wiley: Chichester, England,
1999.

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
ACTIVE BASE ISOLATION OF BUILDINGS SUBJECTED TO SEISMIC EXCITATIONS 1511

5. Ramallo JC, Johnson EA, Spencer Jr BF. Smart base isolation system. Journal of Engineering Mechanics 2002;
128(11):1088–1099.
6. Buckle I, Nagarajaiah S, Ferrell K. Stability of elastomeric isolation bearings: experimental study. Journal of
Structural Engineering 2002; 128(1):3–11.
7. Kelly JM. The role of damping in seismic isolation. Earthquake Engineering and Structural Dynamics 1999;
28:3–20.
8. Nagarajaiah S, Ferrell K. Stability of elastomeric seismic isolation beargins. Journal of Structural Engineering
1999; 125(9):946–954.
9. International Conference of Building Officials (ICBO). Uniform Building Code, Earthquake Regulations for
Seismic-isolated Structures, vol. 2. Appendix Chapter 16, Whittier, CA, 1997.
10. Yoshioka H, Ramallo JC, Spencer Jr BF. Smart base isolation strategies employing magnetorheological dampers.
Journal of Engineering Mechanics 2002; 128(5):540–551.
11. Reinhorn A, Riley MA. Control of bridge vibrations with hybrid devices. Proceedings of the First World
Conference on Structural Control, Los Angeles, CA, 1994; 50–59.
12. Inaudi JA, Kelly JM. Active isolation. U.S. National Workshop on Structural Control Research, Los Angeles,
1990; 125–130.
13. Nagarajaiah S, Riley MA, Reinhorn A. Control of sliding-isolated bridge with absolute acceleration feedback.
Journal of Engineering Mechanics 1993; 119(11):2317–2332.
14. Yang JN, Wu JC, Reinhorn AM, Riley M. Control of sliding-isolated buildings using sliding-mode control.
Journal of Structural Engineering 1996; 122(2):179–186.
15. Dyke SJ, Spencer Jr BF, Quast P, Sain MK. The role of control-structure interaction in protective system design.
Journal of Engineering Mechanics 1995; 121(2):322–338.
16. Spencer Jr BF, Johnson EA, Ramallo JC. Smart isolation for seismic control. JSME International Journal Series
C 2000; 43(3):704–711.
17. Makris N. Rigidity-plasticity-viscosity: can electrorheological dampers protect base-isolated structures from near-
source earthquakes. Earthquake Engineering and Structural Dyanmics 1997; 26:571–591.
18. Nagarajaiah S, Sahasrabudhe S, Iyer R. Seismic response of sliding isolated bridges with MR dampers. Proceedings
of the America Control Conference, Chicago, IL, vol. 6, 2000; 4437–4441.
19. Shook D, Lin PY, Lin TK, Roschke PN. A comparative study in the semi-active control of isolated structures.
Smart Materials and Structures 2007; 16:1433–1466.
20. Sahasrabudhe S, Nagarajaiah S, Hard C. Experimental study of sliding isolated buildings with smart dampers
subjected to near-source ground motions. Engineering Mechanics Conference Proceedings, ASCE, EM, Austin,
TX, 2000 (CD-Rom).
21. Spencer Jr BF, Dyke SJ, Deoskar HS. Benchmark problems in structural control—Part I: active mass driver
system. Earthquake Engineering and Structural Dynamics 1998; 27:1127–1139.
22. Soong TT, Grigoriu M. Random Vibration of Mechanical and Structural Systems. Prentice-Hall: Englewood
Cliffs, NJ, 1993.
23. Dyke SJ, Spencer Jr BF, Quast P, Kaspari Jr DC, Sain MK. Implementation of an active mass driver using
acceleration feedback control. Microcomputers in Civil Engineering: Special Issue on Active and Hybrid Structural
Control 1994; 11:305–323.
24. Dyke SJ, Spencer BF Jr, Quast P, Sain MK, Kaspari Jr DC. Experimental verification of acceleration feedback
control strategies for an active tendon system. National Center for Earthquake Engineering Research Technical
Report NCEER-94-0024, 1994.
25. Spencer Jr BF, Suhardjo J, Sain MK. Frequency domain optimal control strategies for aseismic protection. Journal
of Engineering Mechanics 1994; 120(1):135–139.
26. WorkSafe Technologies, Valencia, CA, 2009. Available from: http://www.worksafetech.com/.
27. Kasalanati A, Reinhorn A, Constantinou M, Sanders D. Experimental study of ball-in-cone isolation system.
Building to Last: Proceedings of the 15th Structures Congress. ASCE: New York, 1997; 1191–1195.
28. Moog, East Aurora, NY, 2010. Available from: http://www.moog.com/, 1951-present.
29. Shore Western Manufacturing, Monrovia, CA, 2009. Available from: http://www.shorewestern.com/.
30. Quast P, Spencer Jr BF, Sain MK, Dyke SJ. Microcomputer implementation of digital control strategies for
structural response reduction. Microcomputers in Divisional Engineering 1995; 10:13–25.
31. Dyke SJ, Spencer Jr BF, Quast P, Sain MK, Kaspari Jr DC, Soong TT. Acceleration feedback control of MDOF
structures. Journal of Engineering Mechanics 1996; 122(9):907–918.
32. van Overschee P, de Moore B. N4SID: subspace algorithms for the identification of combined deterministic-
stochastic systems. IEEE Automatica 1994; 30:75–93.

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe
1512 C.-M. CHANG AND B. F. SPENCER JR

33. Bayard DS. High-order multivariable transfer function curve fitting: algorithms, sparse matrix method, and
experimental results. IEEE Automatica 1994; 30(9):1439–1444.
34. Auweraer HVD, Guillaume P, Verboven P, Vanlanduit P. Application of a fast-stabilizing frequency domain
parameter estimation method. Journal of Dynamic Systems, Measurement and Control 2001; 123:651–658.
35. Kim SB, Spencer Jr BF, Yun CB. Frequency domain identification of multi-input, multi-output systems considering
physical relationships between measured variables. Journal of Mechanical Engineering 2005; 131(5):461–472.
36. Dickinson BW, Morf M, Kailath T. Minimal realization algorithm for matrix sequences. IEEE Transactions on
Automatic Control 1974; AC19(1):31–38.
37. Ober RJ. Balanced parameterization of classes of linear-systems. SIAM Journal on Control and Optimization
1991; 29(6):1251–1287.
38. Chang CM, Wang Z, Spencer Jr BF. Application of active base isolation control. Proceedings of SPIE,
San Diego, CA, vol. 7292, 2009; 72923901–72923912.
39. Chen CT. Linear System Theory and Design. Oxford University Press, Inc.: New York, 1998.
40. Narasimhan S, Nagarajaiah S, Johnson EA, Gavin HP. Smart base-isolated benchmark building. Part I: problem
definition. Journal of Structural Control and Health Monitoring 2006; 13(2–3):573–588.

Copyright q 2010 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2010; 39:1493–1512
DOI: 10.1002/eqe

You might also like