Sie sind auf Seite 1von 15

Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

Contents lists available at ScienceDirect

Journal of Wind Engineering


and Industrial Aerodynamics
journal homepage: www.elsevier.com/locate/jweia

Wind-induced responses and dynamics characteristics of an asymmetrical MARK


base-isolated building observed during typhoons

Dionysius M. Siringoringo , Yozo Fujino
Institute of Advanced Sciences, Yokohama National University, 79-1 Tokiwadai, Hodogaya-ku, Yokohama 240-8501, Japan

A R T I C L E I N F O A BS T RAC T

Keywords: Responses of an asymmetrical base-isolated building during the passage of Typhoon Roke 2011, Typhoon
Base-isolated building Guchol 2012 and a bomb cyclone strong-wind event on April 3, 2012 were investigated in this paper. Wind-
Typhoon-induced full-scale measurement induced accelerations of the building were recorded by dense array permanent monitoring system and analyzed
Building vibration in detailed by spectra analysis, time-frequency wavelet analysis and system identification. Results of analysis
Typhoon response
revealed that wind-induced responses of the buildings were dominated by the first mode which is the
System identification
fundamental flexural mode coupled with torsion of the upper stories. Building responses demonstrated non-
linear characteristics of natural frequencies, damping ratio and mode-shapes. Natural frequencies of funda-
mental modes decreased with the increase of acceleration amplitude, with the maximum reduction of 8–10% of
natural frequencies observed during the peak wind speed. Damping estimates of the fundamental modes
increased with the increase of accelerations, while characteristics of mode-shapes were also found to be
dependent on the wind speed. Further analysis revealed that non-linear characteristics of building responses are
related to isolator deformations caused by wind-induced static force.

1. Introduction base-isolated building (Japan Society of Seismic Isolation JSSI, 2009).


Despite the efforts to investigate wind-response of base-isolated
Base isolation is a popular seismic mitigation technique for build- building through modeling and simulation, there have been very few
ings and has been applied to more than 150 tall buildings with the studies that are based on real observation of full-scale base-isolated
height of 60 m or more (Takenaka et al., 2004). As the number of base- building especially due to strong wind excitation (Yasui et al., 2002;
isolated tall buildings increases, consideration for wind-resistant de- Sato et al., 2012). As comparison, there have been numerous studies on
sign has become more significant than before. Unlike earthquake, wind the wind-response of tall conventional fixed base buildings under
load has two different main characteristics, namely, long duration of strong wind events such typhoons (Campbell et al., 2005; Li et al.,
excitation and having static component (mean force). Because of 2000, 2003). Based on these studies, several conclusions are obtained
isolator characteristics, natural period of a base-isolated building is such as the amplitude-dependent characteristics of damping and
longer than ordinary building resulting in larger wind-induced dis- natural frequency (Li et al., 2000, 2003), and comparison between
placement of the upper story. predicted and measured wind-induced response of conventional fixed
There have been some studies on the behavior of base-isolated base tall buildings (Jeary and Ellis, 1983). Recently, more base-isolated
buildings subjected to wind excitation. Most of them focus on modeling buildings have been instrumented with permanent monitoring system
building response to wind excitation (Kareem, 1997; Chen and to observe their performance during large earthquake. The monitoring
Ahmadi, 1992; Henderson and Novak, 1989), addressing the service- system can also be used to capture the wind-induced responses during
ability aspect of wind-induced vibration (Liang et al., 2002) and strong wind or typhoon.
comparing wind-induced response and earthquake response of base- The objective of this study is to investigate wind-induced responses
isolated building (Vulcano, 1998). Several studies have experimentally and dynamic characteristics of a base-isolated building from full-scale
investigated behavior of isolator component especially creep behavior structural responses recorded during the passages of typhoons. At first
that becomes an important consideration for relatively long duration of the paper outlines measurement program and describes the main
wind excitation (Takenaka et al., 2004, 2009; Kochiyama et al., 2004). characteristics of typhoon-induced building responses in detailed using
Result of the study is included in the design guideline for wind effect of temporal, spectral and time-frequency wavelet analysis. Further,


Corresponding author.
E-mail addresses: dion@ynu.ac.jp (D.M. Siringoringo), fujino@ynu.ac.jp (Y. Fujino).

http://dx.doi.org/10.1016/j.jweia.2017.04.020
Received 22 July 2016; Received in revised form 26 February 2017; Accepted 26 April 2017
0167-6105/ © 2017 Elsevier Ltd. All rights reserved.
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

Fig. 1. (a) Photo of SIT Building, (b) dimension, orientation and building's plan view with layout of isolators at basement level, (c) Layout and location of monitoring system (note: NRB:
natural rubber bearing, SB: sliding bearing, LD: lead damper and UD: U-shaped steel dampers).

dynamic characteristics of the building obtained from system identifi- 26 units of sliding bearing (SB), 28 units of lead damper (LD) and 33
cations are described with respect to wind speed and direction. Finally, units of U-shaped steel dampers (UD) (Fig. 1). Detail explanation of the
the paper provides discussions on interpretation of the results and isolation system is given elsewhere (Siringoringo and Fujino, 2015a).
compares the results with design condition and observations from A permanent monitoring system consisting of 21 triaxial acceler-
seismic records. ometers and four displacement-meters is deployed on the building
since 2010 as shown in Fig. 1(c). The sensors nomenclature is
2. Description of base-isolated building and monitoring organized as follow, the first index denotes sensor type (A: acceler-
system ometer and D: displacement-meter), the second index represents the
building (M: main and A: annex building), the third index represents
The object structure is a mid-rise asymmetrical base-isolated story level, and the last index shows location of the sensor in the floor
building named SIT Building located in Tokyo-Bay area. The building (i.e. S: South, E: East and W: West). Horizontal accelerations were
consists of two parts: fourteen-story main building (M) and seven-story oriented in the x and y directions per building as opposed to the
annex building (A) (Fig. 1). Both buildings are of braced steel frames commonly used North-South (NS) and East-West (EW) directions. The
and connected at the corner by elevator shaft forming an L-shaped x-axis coincides with the main building's weak axis, while the y-axis
asymmetric structure. Dimension of the main building is 97.2 m long, coincides with the main building's strong axis. To avoid confusion, the
43.2 m wide and 67.5 m high with 19 bays in longitudinal direction and terms x-axis and y-axis according to Fig. 1. will be used throughout the
6 bays in transverse direction. Annex building is 81 m long, 21.6 m paper.
wide and 31.2 m high with 16 bays in longitudinal direction and 3 bays Eighteen accelerometers are placed on the building and three on the
in transverse direction. The main building has vertical opening in the ground outside the building. It should be mentioned, however, that
middle, starting from the second to the seventh floor, which divides the displacement-meters that measure relative displacement between base-
building into the west section (MW) and east section (ME). Concrete ment and the first floor are activated by an underground trigger that
slab connects both buildings at the basement level and on top of the functions only when ground motion with acceleration larger than 5 cm/
slab the isolation systems are placed. s2 occurs as in the case of earthquake. These sensors were not triggered
The main building's orientation and axis direction are depicted in by typhoon-induced vibration; therefore, in this paper we present
Fig. 1.(b). The building orientation is such that the x-axis deviates 42° analysis only from acceleration responses of upper stories. The accel-
clockwise from the north. Due to its unique shape, the building has erometers are the servo type with resolution of 0.01 cm/s2 and
significant eccentricity. The eccentricity ratio is defined as the distance measurement range ± 2000 cm/s2. Sensor network is connected to a
between center of stiffness and center of mass with respect to building GPS controller unit to provide global reference position and to
length in the corresponding direction. The ratio is presented separately synchronize time recording with 4 ms maximum delay time among
for each floor and each building section in x- and y-directions. The sensors (Siringoringo and Fujino, 2015a). The monitoring system
eccentricity in y-direction is generally larger than in x-direction for records building accelerations at 100 Hz sampling frequency and stores
annex and main building; and the eccentricity ratios in some floors are the data in a server for further analysis.
larger than 10%. Wind records were provided by Edogawa weather station, the
The building is base isolated with total of 146 units of isolator and closest weather station located approximately 2 km southeast of the
damper system consisting of 59 units of natural rubber bearing (NRB), building, since onsite wind measurement system was unavailable on

184
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

Fig. 2. Ten-minute averaged and instantaneous peak wind speed and average wind direction for the case of (a) Typhoon Roke, September 5, 2011 (24 h), (b) Bomb-cyclone, April 3,
2002 (24 h), (c) Typhoon Guchol, June 19–20, 2012 (48 h). Note: arrows in the building inset denote the dominant wind direction with respect to building's orientation.

the building at the time of the typhoons and storm event. Considering direction. Variations of 10-min mean and maximum instantaneous
topography of the area surrounding the building and the proximity of wind speed during the storm are shown in Fig. 2(b).
Edogawa weather station, the collected wind information is deemed
reasonable to reflect overall wind characteristics affecting the building. 4. Characteristics of building's dynamic responses
The wind records consist of 10-min average wind speed and direction,
as well as maximum instantaneous wind speed within ten-minute 4.1. Response amplitude with respect to wind speed
duration. The wind data was recorded at the standard height of 10.2 m
from sea level (Japan Meteorological Agency (JMA) (2016)). Acceleration responses recorded at the east corner of the 14th floor
of main building (sensor AM14E) are shown in Fig. 3 for typhoon Roke,
3. Description of typhoons and strong wind event Guchol and bomb cyclone. The records were obtained during 10-min
duration of the largest wind speed of the respective events. For the
In this study, the building responses were recorded during two typhoon Roke 2011 the peak acceleration recorded by sensor AM14E
typhoons and one low pressure storm event called bomb cyclone. The first was 4.74 cm/s2 at 18:29 Japan Standard Time (JST) of Sept 5, 2011
typhoon is the typhoon Roke 2011 or the 15th typhoon of 2011 in Japanese caused by the maximum instantaneous wind speed of 41 m/s. The peak
typhoon counting system that passed through the building location on acceleration due to bomb-cyclone was about 2.31 cm/s2 occurring at
September 5th, 2011. The typhoon recorded the lowest instantaneous the instantaneous time of the maximum wind speed 33.4 m/s. In the
mean sea-level pressure and the maximum recorded wind speed from all case of typhoon Guchol, the typhoon gained strength rather slowly but
meteorological stations are 940 hPa and 43.7 m/s, respectively. At lasted for two days. The increase in building acceleration was noted at
Edogawa weather station, the maximum recorded instantaneous wind 19:00 JST in the first day June 19, 2012 and reached the maximum
speed was 41 m/s (Fig. 2.a). The dominant wind direction during the around midnight with the peak building acceleration of 3.73 cm/s2 due
largest wind speed of typhoon Roke was between southwest and southeast to maximum wind speed of 32.6 m/s.
direction which was nearly parallel to the building's x-axis. Variations of 10-min root-mean square (RMS) and peak building
The second typhoon is Typhoon Guchol 2012 or the 4th Typhoon of accelerations during the 10-min for the entire day of typhoon are presented
2012 that passed through area near the building on June 19, 2012 in Fig. 4. The largest accelerations for the three events were recorded by
afternoon until June 20 evening. The lowest instantaneous mean sea- sensor AM14E in x-direction, and the amplitude varied between 4.5 and
level pressure and maximum recorded speed of the typhoon from all 6 cm/s2. In y-direction, the peak accelerations were smaller between 2 and
meteorological stations are 930 hPa and 51 m/s, respectively. Typhoon 4 cm/s2. By comparing the daily variation of wind speed as shown in Fig. 2
Guchol recorded the maximum wind speed 32.6 m/s at Edogawa and the building acceleration, one can observe that as expected, building
weather station, with dominant winds direction were from south and accelerations increase with increasing wind speed. It should be mentioned,
southeast that was almost perpendicular to building's y-axis. Fig. 2(c) however that there is a sudden increase in building acceleration around
illustrates the 10-min mean and maximum instantaneous wind speed 21:15 JST of typhoon Roke. During this time instance, the peak and RMS
for typhoon Guchol. of accelerations are very different with the ones that precede and follow
The third event is a wind storm caused by an extra-tropical cyclone after, whereas the wind speed at this time instant is smaller than before.
also known as bomb cyclone. The storm occurred on April 3, 2012 with This spike in acceleration was not caused by wind response but due to a
the recorded maximum instantaneous wind speed of 51.1 m/s at small earthquake Mw 4.3 about 200 km to the northeast side of the
Tobishima Sakata, Yamagata Prefecture. The highest measured wind building.
speed at the Edogawa weather station was 33.4 m/s, with the dominant Fig. 5 illustrates the relationships between the peak (Pacc) and RMS
wind direction during the storm was between southwest to south values (σacc) of 10-min x and y-direction acceleration response with respect

185
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

5
Acc (cm/s ) 2
Peak Acceleration : 4.74 cm/s
2

-5
(a) 18:20 18:21 18:22 18:23 18:24 18:25 18:26 18:27 18:28 18:29 18:30

5
Acc (cm/s )
2

-5
18:20 18:21 18:22 18:23 18:24 18:25 18:26 18:27 18:28 18:29 18:30
(b)

2
Acc (cm/s )
2

-2 2
Peak Acceleration : 2.31cm/s
(c) 19:40 19:41 19:42 19:43 19:44 19:45 19:46 19:47 19:48 19:49 19:50

2
Acc (cm/s )
2

-2
19:40 19:41 19:42 19:43 19:44 19:45 19:46 19:47 19:48 19:49 19:50
(d)

5
Acc (cm/s )
2

0
2
Peak Acceleration : 3.73cm/s
-5
00:00 00:01 00:02 00:03 00:04 00:05 00:06 00:07 00:08 00:09 00:10
(e)

5
Acc (cm/s )
2

-5
00:00 00:01 00:02 00:03 00:04 00:05 00:06 00:07 00:08 00:09 00:10
(f) Time (Hour)

Fig. 3. Acceleration responses recorded on the top of main building (AM14E) for: (a) typhoon Roke X-direction, (b) typhoon Roke Y-direction, (c) bomb-cyclone X-direction, (d) bomb-
cyclone Y-direction, (e) typhoon Guchol X-direction, (f) typhoon Guchol Y-direction.

to 10-min mean wind speed for all three events. It is evident from this The figures with the constant of regression values of a1, a2, b1 and b2
figure that the increases in acceleration responses were associated with the listed in Table 1 demonstrate that accelerations in x-direction that is the
increase in the approaching mean wind speed. The regression curves of weak-axis of the main building increase more significantly than accelera-
acceleration responses presented in these figures are the fitted power tions in y-direction. Note that the peak accelerations on the main building
regression curve with 95% confidence bounds and expressed as: in x and y directions are due to south to southwest winds, which is the
dominant wind directions for the three events. Looking at the wind
σacc = a1 U b1 (1)
directions and building accelerations as shown by the values of coefficient
of regression during the peak of typhoons, it can be concluded that
Pacc = a2 U b2 (2)
typhoon-induced responses of the main building were mainly the along-
where the values of a1, a2, b1 and b2 as the results of curve fitting are given wind responses.
in Table 1.

Fig. 4. Ten-minute averaged and instantaneous peak of floor acceleration of the main building (14th Floor) for the case of (a) Typhoon Roke, September 5, 2011, (b) Bomb-cyclone,
April 3, 2002, (c) Typhoon Guchol, June 19–20, 2012.

186
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

Fig. 5. Relationship between 10-min mean wind velocity and 10-min mean and peak building accelerations recorded at the 14th floor of main building in X direction (AM13E-X) for: (a)
Typhoon Roke, (b) Bomb-cyclone, (c) Typhoon Guchol; and in Y direction (AM13E-Y) for (d) Typhoon Roke, (e) Bomb-cyclone, (f) Typhoon Guchol.

Table 1
Parameters of regression curved fitted in Eqs. (1) and (2). artificially offset to show that peaks of all spectra located in the same
frequency. The figures reveal clearly the dominant influence of low
Typhoon Sensor Root mean squared of Peak acceleration (Pacc) frequency peak at 0.58 Hz for the east corner of the main and annex
acceleration (σacc)
building in both x and y directions. Whereas for the west corner of
a1 b1 R2 a2 b2 R2 main building that is the middle corner of the L- shape building
frequency peak at 0.74 Hz dominates the spectra in x direction. Also
Roke AM14E-X 3.70E 2.875 0.8893 6.61E−05 3.259 0.9035 note that the same pattern of spectra peaks appear on different floors
−05
located on the same side indicating continuity of the responses among
AM14E-Y 1.89E 2.145 0.815 2.81E−05 3.344 0.8605
−04
the floors. As expected, the height of spectra peaks that indicate
Bomb AM14E-X 5.09E 2.047 0.8178 8.93E−04 2.472 0.7892 vibration level are different according to the floor in that the higher
Cyclo- −04 floor has higher level of vibration and thus higher peak of spectra. The
ne AM14E-Y 2.38E 1.292 0.7773 4.37E−03 1.717 0.6745 fact that spectra peak at 0.58 Hz and 0.78 Hz appear in both x and y
−03
directions indicate the occurrence of torsional vibration modes. We
Guchol AM14E-X 2.30E 2.352 0.7671 3.59E−04 2.783 0.7342
−04 shall see this in more detail when discussing the mode-shapes from
AM14E-Y 1.68E 1.452 0.6394 1.60E−03 2.079 0.6474 system identification.
−03 Fig. 7 illustrates the spectra of building accelerations for 10-min
period at 10:30 JST earlier that day where RMS of wind speed was
between 2.3 m/s and the peak was 4.1 m/s, as a comparison of spectral
4.2. Spectral characteristics of building accelerations
characteristics during moderate or ambient wind. The figures exhibit
similar pattern with that of strong wind accelerations. The spectra are
Most of the records consist of small to moderate accelerations
dominated by low frequencies peaks while contribution of higher
except for several hours when the typhoon peaks passed the building
resonant frequencies at 2 Hz, 3 Hz and 3.5 Hz are more apparent this
location. For typhoon Roke the large accelerations period, which is
time as shown by the relatively higher peaks for those respective
when peak accelerations on top of the building are larger than 0.5 cm/
frequencies. It is interesting to note that frequencies associated with
s2, lasted for about 6 h from 16:00 to 22:00 JST. For bomb-cyclone, the
the peaks for the ambient wind are slightly higher than that the
large acceleration period was about 8 h from 15:00 to 23:00 JST. And
corresponding peaks during the strong wind. For example, frequency of
for the typhoon Guchol the duration of large accelerations period was
the first peak changed to 0.63 Hz from 0.58 Hz, while the second peak
about 9 h, starting from June 19, 2012 at 19:00 JST until 04:00 JST
shifted to 0.79 Hz from 0.7 Hz.
the next day.
The same trend of frequency shift was observed from building
Spectral characteristics of building accelerations are evaluated for
accelerations due to the bomb-cyclone and typhoon Guchol as listed in
moderate and strong wind excitations to show the difference in
Table 2. In both cases, the acceleration spectra during the highest wind
characteristics. Fig. 6 shows the spectra amplitude of accelerations of
speed are characterized by lower resonant frequency peaks at the range
the main and annex building at 18:30 JST for the 10-min period of the
of 0.58–0.6 Hz for the first peak and 0.73–077 Hz for the second peak.
strongest wind of typhoon Roke (i.e. wind speed RMS=30.1 m/s and
Meanwhile for lower wind speed ( < 3.4 m/s) the peaks of acceleration
peak=41 m/s). Note that in the figures the x-axis of each spectrum was
spectra are at the higher frequency range of 0.64–0.65 Hz for the first

187
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

AM14E-X AM09E-X AM07E-X AM04E-X AM14E-Y AM09E-Y AM07E-Y AM04E-Y


15000 10000
Fourier Amplitude
0.58Hz 0.58Hz
0.74Hz
10000
5000
5000

(a) 0 (b) 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
AM14W-X AM09W-X AM07W-X AM04W-X AM14W-Y AM09W-Y AM07W-Y AM04W-Y
10000 10000
Fourier Amplitude

0.58Hz
0.74Hz

5000 5000

(c) 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 (d) 0 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

6000 4000
Fourier Amplitude

0.58Hz AA7S-X AA4S-X 0.58Hz AA7S-Y AA4S-Y


0.74Hz 3000
4000
2000
2000
1000

0 0
(e) 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 (f) 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency (Hz) Frequency (Hz)

Fig. 6. Fourier spectral characteristics of building accelerations during the strongest wind period of typhoon Roke (18:30 JST Sept 5,2011, rms=30.1 m/s and peak=41 m/s) for: (a) east
corner of main building in X-direction, (b) east corner of main building in Y-direction, (c) west corner of main building in X-direction, (d) west corner of main building in Y-direction, (e)
south corner of annex building in X-direction, and (f) south corner of annex building in Y-direction.

peak and 0.78–0.79 Hz for the second peak. The results indicate 4.3. Time-frequency characteristics of building accelerations
reduction of resonant frequencies of the building during strong wind
excitation. We shall see this in more detailed when discussing the The Fourier spectra analysis provides a basic description on the
variation of resonant frequencies with respect to wind speed from energy distribution of acceleration responses in frequency domain for
system identification. stationary signal. In the case of typhoon responses, however, it is not
uncommon that they become non-stationary (see for instance (Xu and

AM14E AM09E AM07E AM04E AM14E AM09E AM07E AM04E


1000 800 0.63Hz
0.63Hz
Fourier Amplitude

0.79Hz 600

500 400

200

(a) 00 (b) 0
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

AM14W AM09W AM07W AM04W AM14W AM09W AM07W AM04W


1000 0.79Hz 800
Fourier Amplitude

0.63Hz
600

500 400

200

0 0
(c) 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 (d) 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

400 300
0.63Hz
Fourier Amplitude

AM07E 0.63Hz AM07E


300 0.79Hz AM04E AM04E
200
200
100
100

0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
(e) (f)
Frequency (Hz) Frequency (Hz)

Fig. 7. Fourier spectral characteristics of building accelerations during the low wind period of typhoon Roke (10:30 JST Sept 5,2011, rms=2.3 m/s and peak=4.1 m/s) for: (a) east
corner of main building in X-direction, (b) east corner of main building in Y-direction, (c) west corner of main building in X-direction, (d) west corner of main building in Y-direction, (e)
south corner of annex building in X-direction, and (f) south corner of annex building in Y-direction.

188
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

Table 2 peaks for 5-h scalogram from 17:00 to 22:00 JST are lower than 0.6 Hz
Peaks resonant frequencies on spectra of building accelerations identified during (Fig. 8(b)). The results agree well with the spectral characteristics as
typhoons at peak wind speed and low wind speed (ambient).
previously described by Figs. 6 and 7.
Typhoon roke Bomb-cyclone Typhoon guchol To have a better description on evolution of frequency content, the
wavelet instantaneous spectra analysis is performed by slicing the
Ambient Peak Ambient Peak Ambient Peak scalogram at a specific time and expanding the view in frequency
9:50a 18:30 3:30 19:30 5:50 00:20 domain. The resulting plot shows variation of wavelet coefficients
[0.7 m/s] [41 m/s] [3.4 m/s] [32 m/s] [1.0 m/s] [34 m/s]
across the range of frequencies. Examples of instantaneous wavelet
1st peak 0.65 Hz 0.580 Hz 0.64 Hz 0.599 Hz 0.65 Hz 0.589 Hz spectra for several time instants are shown in Fig. 8(c). Again from
2nd peak 0.79 Hz 0.740 Hz 0.78 Hz 0.761 Hz 0.79 Hz 0.760 Hz these figures, one can observe clearly a general decreasing trend in the
peak frequency as wind speed increases. During moderate wind speed
Note:
a as shown in time instant between 14:30 and 16:30 JST, the peak
Time is in Japan Standard Time (JST).
frequency is about 0.62 Hz and 0.613 Hz respectively. When the wind
speed as well as building acceleration reached the peak around 18:00–
Chen, 2004; Chen and Xu, 2004)), which means characteristics of
19:00 JST, the peak frequency dropped to about 0.56 Hz. When wind
energy distribution and frequency vary with respect to time. Under this
speed and building acceleration decreased near the end of the day, the
condition, it is necessary to use signal processing tools that can produce
peak frequency recovered back to 0.62 Hz. The same trend was
a time-frequency distribution of the responses. For this purpose, the
observed from the bomb-cyclone and typhoon Guchol data (Fig. 9).
Morlet-type Wavelet transform is utilized. The energy distribution in
For the bomb-cyclone, the instantaneous spectra of acceleration show a
the Morlet-type Wavelet transform is represented by plotting the
drop in frequency from about 0.62 Hz in the beginning of the event to
squared modulus of wavelet coefficients C(a,t) called the scalogram
0.55 Hz during the peak wind speed at 21:30 JST. The peak frequency
plot (Cohen, 1995).
was later recovered when the strong wind period end around midnight.
An example of scalogram of acceleration from sensor AM14E-X for
Similarly, for the typhoon Guchol, the peak frequency began to drop
the typhoon Roke is illustrated in Fig. 8. The energy distribution in
when the typhoon gained its strength around 23:00 JST and reached
time and frequency domain can be seen clearly by the contrast in
the lowest frequency at 0.58 Hz around midnight. The peak frequency
colors. It is evident from the figures that low order frequencies that
was later recovered to 0.62 Hz after the high wind speed period was
correspond to fundamental modes (i.e the first and second modes)
over.
within the range of 0.58–0.8 Hz dominate the energy distribution
throughout the entire acceleration responses. Higher order modes at
about 2 Hz and 3.5 Hz were also observed, but their contributions are 5. System identification methodology and results
rather weak. The figures show that for small to moderate wind speed,
as shown by the 5-h scalogram from 12:00 to 17:00 JST, the dominant In order to extract modal properties of the building from recorded
peaks are slightly higher than 0.6 Hz. Meanwhile, the corresponding dynamic responses, two system identification methodologies were

Fig. 8. Morlet Wavelet Scalogram of acceleration responses recorded by sensor AM14E in X direction, showing (a) 5-h of time-frequency variation during small wind excitation, (b) 5-h
of time-frequency variation during the peak wind speed, (c) time evolution of wavelet instantenous spectra.

189
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

0.015 0.08 0.2 0.08


t=20:20 0.8 t=21:30s
t=16:30 t=18:00
2

0.06 0.15 t=23:10


Wavelet Coeff

0.06
0.01 0.6
f=0.618Hz 0.04 0.1 0.04
f=0.613Hz 0.4 f=0.625Hz
f=0.571Hz f=0.55Hz
0.005
0.02 0.05 0.2 0.02

0 0 0 0 0
(a) 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Freq(Hz) Freq(Hz) Freq(Hz) Freq(Hz) Freq(Hz)
0.04 0.4 4 0.5 0.05
t=23:10 t=01:30s t=03:30
t=21:30 t=00:30
Wavelet Coeff 2

0.03 0.3 0.4 0.04


3
0.3 0.03
0.02 0.2 2
f=0.64Hz f=0.62Hz 0.2 0.02 f=0.621Hz
f=0.58Hz f=0.59Hz
0.01 0.1 1 0.1 0.01
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0
(b) 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Freq(Hz) Freq(Hz) Freq(Hz)
Freq(Hz) Freq(Hz)

Fig. 9. Time evolution of Morlet Wavelet instantenous spectra of acceleration responses recorded by sensor AM14E in X direction, for (a) bomb-cyclone and (b) typhoon Guchol.

utilized, namely, the multi-output time-invariant system identification throughout the entire 10-min records. Therefore, a piece-wise linear
using information matrix (SRIM (Juang, 1997)) and the covariance- analysis was conducted using shorter moving time window, during
driven stochastic subspace identification (SSI (Peter van Overschee and which the modal parameters were assumed to remain constant. To this
Bart De Moor, 1996)). The SRIM method is based on the realization of end, the total 10-min acceleration responses were divided into ten time
dynamical system from cross-correlation functions of multi-channel windows consisting of 1-min data each from which a set of modal
accelerations. The identification process starts by constructing the parameters is generated.
multi-output correlation matrix defined as a correlation matrix be- Table 3 list the representative values of frequencies and damping
tween acceleration responses from multiple channels on the building to ratios obtained from both system identifications. Two cases are selected
a specific acceleration response, so called the reference channel. In this for each typhoon, namely the frequencies and damping during peak
case the acceleration from channel AM0W (x and y directions) were wind excitation and smaller wind which can be considered as the case
selected as the reference. Further details on description of the SRIM for ambient excitation. Frequencies of both vibration modes identified
method and implementation on the multi-channel output-only system from SRIM and SSI are very similar indicating the accuracy of
are presented elsewhere (Siringoringo and Fujino, 2008). Meanwhile identification, and their values are also very close to those obtained
the SSI method is based on covariance matrix computed from from spectra of accelerations. One can observe from the table that the
covariance of all channels on the upper-stories in x and y directions. identified frequencies of the first and second modes drop during the
Details on description of the SSI method on the multi-channel output highest wind speed. The results that are very similar with the peak
scheme can be found elsewhere (Peter van Overschee and Bart De resonances observed from frequency spectra of accelerations. The table
Moor, 1996). also show that although not very significant, the damping ratios
Both system identifications are time domain and based on linear obtained during high wind speed are generally larger than that
system assumption, which means modal parameters are considered to estimated from low wind responses. The tendency is consistent for
remain constant within the specific time-window of analyzed data. both modes from two system identifications.
However, considering the possibility of non-linear responses of the The table provides information on frequencies and damping
building during large excitation, this assumption may not be satisfied estimated from earthquakes, that is the biggest earthquake March 11,

Table 3
Examples of Identified natural frequencies and damping ratio for lower wind speed (ambient) and at the peak wind speed during typhoons and comparison with frequencies estimated
from FEM and earthquakes.

Typhoon roke Bomb-cyclone Typhoon Guchol

Ambient Peak wind speed Ambient Peak Wind speed Ambient Peak wind speed

9:50 [0.7 m/s]a 18:30 [41 m/s] 3:30 [3.4 m/s] 19:10 [32 m/s] 5:50 [1 m/s] 24:20 [34 m/s]

Frequency (Hz)**

1st Mode 0.620/0.630 0.556 /0.586 0.643/0.631 0.566/0.571 0.640/0.645 0.581/0.585


2nd Mode 0.782/0.784 0.740/0.749 0.784/0.773 0.755/0.761 0.784/0.793 0.760/0.756

Damping ratio (%)


1st Mode 1.54/2.18 3.07/3.06 1.80/2.26 3.06/2.54 1.51 /1.67 2.92/2.35
2nd Mode 1.94/1.10 2.77/2.62 1.67/2.31 3.44/2.84 1.16 /1.22 2.18/2.54

Finite Element Model Freq (Hz) 2011 March 11 Earthquake


Assumed Initial Stiffness of Isolator Assumed Large isolator deformation Main shock Aftershock(Eq. (4))
Freq (Hz) Damp.Ratio (%) Freq (Hz) Damp.Ratio (%)
1st Mode 0.50 0.16–0.18 0.440 14.60 0.530 9.780
2nd Mode 0.64 – 0.580 7.22 0.700 3.830

Note:
a
Frequency and damping ratio were identified from system identifications and presented as SRIM/SSI.

190
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

2011 and Aftershock (Eq. (4)) (Siringoringo and Fujino, 2015a, 2015b) the results from SSI system identifications are slightly higher than that
for comparison. During the largest earthquake, the frequency dropped of SRIM for the same interval of wind responses, but overall values are
significantly to 0.44 Hz and 0.58 Hz for the first and second mode, very similar. The general trend is that the natural frequency estimates
respectively. The significant frequency drop is consequence of func- decrease with the increase of RMS of building accelerations.
tioning of isolation system and this is accompanied by increase in
damping. In the case of aftershock, the frequency is also smaller 5.2. Variation of identified damping ratios with respect to building
compared to structure frequencies during the ambient excitation, but responses
not as significant as the main shock. The natural frequencies estimated
during aftershock are comparable with the values obtained from System identifications that are based on the realization of dynami-
typhoon during the peak response. Comparison between wind-induced cal system such as the SRIM and SSI methods are known to give
and earthquake-induced frequencies of the building will be discussed in relatively accurate estimates of natural frequency and mode-shapes.
more detail in later part of this paper. Estimation of accurate damping ratio is more difficult and highly
subjected to factorization and selection of system order in addition to
5.1. Variation of identified natural frequencies with respect to condition of the signal. Variations of damping estimates are generally
building responses larger, especially for short duration of the records or when the system is
subjected to narrowband nonstationary excitation (Siringoringo and
In general, both system identifications generate three dominant Fujino, 2008).
modes below 1.00 Hz. Comparisons with the previous identification The damping ratios of the first two modes were identified using 1-
from seismic response and finite element model reveal that the modes min acceleration records by implementing the SRIM and SSI identifi-
can be classified as the first translational mode of main building at cations. Generally, the damping ratios were identified between 0.5–5%.
0.55–0.64 Hz, torsional mode of main building at 0.74–0.80 Hz and The largely scatter results, however, create difficulty in drawing a trend
torsional mode of annex building at 0.85–0.95 Hz. In the following or meaningful conclusion directly from 1-min acceleration records with
section, we shall focus the analysis on the first two modes, since they respect to the response amplitude. To give a better view of the results,
are the modes that govern dynamic responses of the building during the average and standard deviation of damping ratios during 10-min
typhoons and bomb-cyclone as shown in acceleration spectra and time- data are computed and the results are shown in the Fig. 12 as the dot
frequency wavelet analysis. and bar, respectively. The results still reveal quite large variation; but
Fig. 10 illustrates the time variation of natural frequencies of the despite the variation there is a general tendency that damping ratios
first two modes generated by both system identifications. Each data increase at the period of largest wind speed. The gradual increase of
point in the figure corresponds to frequency estimates from 1-min damping ratio is noted in the time period between 17:00 and 20:00 JST
acceleration records. Note that both system identifications give very for typhoon Roke, 18:00 and 21:00 JST for bomb-cyclone; 23:00 and
similar results during the observation time indicating high reliability of 2:00 JST for typhoon Guchol. The similar tendency is observed from
frequency estimates. In general, the frequency estimates are relatively both system identification methods as shown in the two columns of
constant throughout the observed period except during the time when figure.
high wind speed occurs. During the period of largest wind speed, To have comparison of damping estimates that are not obtained
gradual frequency reductions appear as noted in the time period from system identification based on realization technique, a separate
between 17:00 and 20:00 JST for typhoon Roke, 18:00 and 21:00 damping identification procedure is conducted. Modal damping ratio is
JST for bomb-cyclone; 23:00 and 2:00 JST for typhoon Guchol. The estimated by computing the autocorrelation of acceleration at each
average frequency reductions are about 10% and 8% for the first and sensor node and applying a band-pass filter to the autospectrum to
second mode, respectively. The general trends of natural frequencies isolate the mode of interest. Afterwards, the single-mode autospectrum
resulted from the two system identifications are shown in Fig. 11 with is transformed to time domain using an inverse Fourier transform to
respect to RMS of accelerations on the top of the building. In general, obtain the free-vibration response. Modal damping ratio is extracted by

Fig. 10. Time variations of the first and second mode natural frequencies identified from system identification SRIM during (a) Typhoon Roke, September 5, 2011, (b) Bomb-cyclone,
April 3, 2002, (c) Typhoon Guchol, June 19–20, 2012. Time variations of the first and second mode natural frequencies identified from system identification SSI during (d) Typhoon
Roke, September 5, 2011, (e) Bomb-cyclone, April 3, 2002, (f) Typhoon Guchol, June 19–20, 2012.

191
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

Fig. 11. Variation of natural frequency of the (a) first mode and (b) second mode identified from 1-min wind-induced accelerations of the building with respect to the RMS of top-floor
acceleration.

Fig. 12. Time variations of damping ratios for the first and second mode identified from system identification SRIM during (a) Typhoon Roke, September 5, 2011, (b) Bomb-cyclone,
April 3, 2002, (c) Typhoon Guchol, June 19–20, 2012. Time variations of damping ratios for the first and second mode identified from system identification SSI during (d) Typhoon
Roke, September 5, 2011, (e) Bomb-cyclone, April 3, 2002, (f) Typhoon Guchol, June 19–20, 2012.

the logarithmic decrement method from the free-vibration response. In respective direction especially during large wind excitation, assumption
principle, the autocorrelation of the response of a single-degree-of- of single-degree-of-freedom vibration can be applied to the autocorre-
freedom system can be considered proportional to its free vibration lation of the acceleration responses. Generally, it is noted that the
response (Brincker et al., 2001; Magalhães et al., 2010). accuracy of damping estimates obtained from free decay response
In the case of acceleration responses of the building, there are two depend on the number of cycles used when extracting the free decay
modes that significantly dominate the response as revealed by spectra response. This suggests that a longer data is essential to give better
figures, that is the first mode from about 0.55–0.62 Hz that can be estimates. For this reason, the damping estimation is conducted using
extracted from sensors in x-direction of east corner of the main 10-min instead of 1-min acceleration data.
building (AME) and the second mode 0.7 Hz that can be extracted Results of damping ratio estimates are plotted in Fig. 13(b) with
from sensors in x-direction of west corner of the main building (AMW). respect to the root-mean-squared of accelerations of top floor. For
Since both modes are well separated and very dominant in its comparison, the average damping ratios for 10-min data obtained from

192
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

Fig. 13. Variation of damping ratio with respect to the RMS of top-floor acceleration. Identified from system identification methods using 10-min wind-induced accelerations (a) first
mode and (b) second mode. Estimated by autocorrelation of top-floor acceleration (c) first mode and (d) second mode. Comparisons with damping ratios estimated from small
earthquakes: (e) first mode and (f) second mode (Siringoringo and Fujino, 2008).

system identifications are also shown in the figure. It is also observed identification are shown in Fig. 14. It should be mentioned that since
that damping ratios obtained from autocorrelation functions are both SRIM and SSI system identification results in the identical mode-
generally larger than that obtained from system identifications for shapes for the same frequency, only one representative mode-shape is
the same time interval. Generally, the damping ratios obtained from shown in the figure for each frequency. In the figure, modal displace-
autocorrelation functions are between 1.5–5% where the first mode's ments of each floor are normalized to the maximum absolute value of
damping is usually larger than the second mode for the same time the three-dimensional modal displacement. The figures were selected
interval. It is evident from the figure that in general, damping ratios to highlight characteristics of corresponding mode-shapes identified
obtained from system identifications and autocorrelation functions from three cases, namely, the peak of typhoon Roke (18:30 JST), the
increase as the building accelerations increase. The increasing trend is small ambient wind condition (11:00 JST) and large earthquake, which
more obvious for damping ratios obtained from autocorrelation func- in this case is the main shock of 2011 Great East Japan earthquake.
tions than from system identification. Note that in the next section, comparisons of mode-shapes from
The increase of damping is known to be associated with the sources moderate earthquakes will also be explained in detailed.
of damping in the building and their contributions to the total damping As illustrated in the Fig. 14, the first mode has characteristic of
when building response increases (Tamura and Suganuma, 1996). In large modal displacement in y-direction especially at the upper stories
the case of a base-isolated building, there are three possible sources of of main building and smaller participation of motion in x-direction.
damping during the wind-induced vibration, namely; the structural The mode exhibits a typical characteristic of flexural mode with small
damping from upper stories, damping from isolators and dissipative participation of torsional motion on the upper stories that is due to
energy devices at the basement, and aerodynamic damping due to asymmetrical characteristics of the building. The first mode identified
building vibration. from wind-excitation differs from the fundamental base-isolation mode
The higher damping estimates at higher building acceleration can identified from large earthquake March 11, 2011. The mode identified
be attributed to the contribution of damping from isolators and from the large earthquake exhibits a sliding mode characteristic in that
dissipative energy devices, while contribution of the aerodynamics modal displacements are concentrated at the isolation layer. The large
damping may not be so significant considering the wind velocity and modal displacement at the isolation layer is commonly observed when
level of acceleration recorded on the top floor, as well as the total height the isolation has fully functioned as in the case of large seismic
of building of 60.4 m. Contributions of damping from isolators and excitation (Siringoringo and Fujino, 2015a, 2015b).
dissipative energy devices are expected since they are activated due to The main difference between the first mode generated by small and
wind-induced deformation of isolators. Structural damping can be large wind is on the amplitude of modal displacement in x and y
approximated from structural responses during ambient vibration direction. In the case of small or ambient wind, modal displacement in
where the effects of damping from isolators and aerodynamic damping y direction is smaller than that of the strong wind. As a result, degree of
are considered small. In such case, the structural dampings are coupling between x and y direction is larger on the mode from ambient
approximated to be 2% and 1.5% for the first and second mode, wind than that of the stronger wind. From previous results of earth-
respectively. quake-induced mode (Siringoringo and Fujino, 2015b), we know that
Fig. 13(c) compares the trend of damping increase with respect to large modal displacement in y-direction at the lower story is associated
increase in building's response for small earthquakes. Two points are with condition of isolator. Similarly, we can relate the difference in
noted from the figures: (1) the trend generally agrees well with the modal displacement during ambient and strong wind with the condi-
estimated from wind-induced response for the same interval of tion of isolator deformation. Therefore, the difference in mode shape
building's accelerations, (2) the estimates of structural damping from amplitude is thought to be attributed to larger deformation of isolators
small earthquakes are found within good agreement with the ones under strong wind excitation. As will be discussed later, this deforma-
obtained from wind-induced responses. tion can be related to the drag force acting on the building whose
magnitude is proportional to the wind speed.
5.3. Variation of identified mode-shapes with respect to building As a torsional mode of main building, the second mode shows large
responses modal displacements concentrated on top corner of the main building
with opposite phase between the east and west corner. Meanwhile,
Typical shapes of the first two modes resulted from system modal displacements of the superstructure of annex building are

193
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

Fig. 14. Estimated mode-shape of the building during the peak acceleration of wind-induced response due to typhoon Roke for (a) the first mode 0.57 Hz and (b) second mode 0.75 Hz.
(Note: bold line=strong wind, dashed line=small wind, dotted lines=large earthquake (March 11, 2011)).

relatively small. Unlike the first mode, the second mode identified from when large deformation occurs on the isolators during large earth-
wind-induced vibration has similar characteristics with the corre- quake, contribution of the torsional modes of upper-structure is
sponding mode identified from large earthquake. This is understand- expected to be insignificant and due to the flexibility of isolation,
able since the torsional modes of the building are not significantly translation mode is expected to dominate vibration of the building.
affected by condition of isolators as revealed in the previously research Frequency reduction and change in fundamental mode shape
(Siringoringo and Fujino, 2015a,b). during strong wind excitation could be explained by employing similar
analogy to earthquake loading especially the effect of shearing force on
isolator deformation. Directions of the oncoming winds during the
6. Relationship between estimated frequency and isolator
largest wind speed period of the three events are on the south-
deformation
southeast or south-southwest (see Fig. 2), indicating the dominant
along-wind excitation with respect to main building's x-axis. In such
The results of spectra analysis, time-frequency wavelet scalogram,
condition, the along-wind exerts drag force on the main building. The
instantaneous wavelet spectra and system identification have all
total shear force (V) acting on the isolator layer is the summation of
established that the fundamental frequencies of the building decrease
shear forces on all floors given by:
with the increase in wind speed. The frequencies, however, recover
n
soon after the wind speeds are back to normal. Since the building does
V MWL = ∑ Fdi
not suffer any structural damage and the upper structure is expected to i =1 (3)
remain elastic throughout the strong wind excitation, the decrease in 1
frequencies is thought to be attributed to the stiffness decrease of the Fdi is the average shear force on each floor defined as Fdi = 2 ρAi Cd Ui 2 ,
seismic isolation layer due to isolators deformation. where ρ, Ai, Cd and Ui are the air density, frontal area, drag coefficient
Previous studies on the building (Siringoringo and Fujino,2015a, and horizontal mean wind velocity of the ith floor, respectively. Since
2015b) revealed that characteristics of fundamental frequencies and the summation of shear forces is computed from mean wind speed, the
associated modes depend on the isolator condition. Since the isolation approach is termed as the mean wind load (MWL) method.
system is designed primarily for seismic loading, two scenarios of For comparison, the total shear forces on the isolators are also
isolator behavior and the effect to modal parameters were considered. computed using the Equivalent Static Wind Load (ESWL). By employ-
One is the case of small and medium earthquake in which the initial ing this approach, the total shear force (V) acting on the isolator layer is
stiffness of isolators is considered, the other one is during large defined as the summation of equivalent static wind loads on all floors:
earthquake, where the isolators enter the secondary stiffness. For the n

small and moderate earthquake, the isolators are expected to remain V ESWL = ∑ WDi
i =1 (4)
on the initial stiffness and the estimated natural frequencies are as
shown in Table 3. In the design for large earthquake (maximum input The ESWL at height i is given as WDi = qi CD GD A, where q i =ρUi2 /2,
larger than 500 cm/s2), substantial deformation of isolation 40 cm is CD is the aerodynamic factor, A the frontal area and GD is the gust
assumed, as a result the first translation mode dominates the building loading factor. The ESWL method used in this comparison utilizes the
vibration. In such condition, the first natural frequency of the building gust loading factor based on base bending moment (BBM) as described
could reach as low as 0.16 Hz and the structure moves as a rigid body in AIJ-RLB-2004 (AIJ-RLB, 2004). The BBM-based gust loading factor
where the main deformation concentrates on isolator layer. In design, is defined as:

194
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

C′g shear force and the estimated isolator displacement resulting from
GD = 1 + gD 1 + φD2 RD
Cg (5) ESWL method are consistently larger than that of MWL. This can be
explained as the consequence of including the gust loading factor,
where, gD is the peak factor, C′g and Cg are the fluctuating and mean whose values are between 2 and 3, when calculating the static load in
coefficient for along-wind overturning moment, ϕD is the mode-shape the ESWL method. The table also includes the maximum isolator
dependent correction factor for along-wind mode and RD the resonance displacements observed from six earthquakes that give the first natural
factor. frequency at the same range with the natural frequency from typhoon
Since the displacement-meters at isolators were not activated events. It is evident from the table that the maximum isolator
during the typhoons, we cannot evaluate directly the measured displacements estimated from MWL method agree well with the values
deformation of isolators due to static wind force. However, we can observed from earthquakes, while those estimated from ESWL are
estimate the deformation by computing the total shear force and use several times larger.
the designed value of lateral stiffness of isolator. To this end, average As suggested in design, the maximum deformation of isolator
shear force or the equivalent static wind load on each floor caused is significantly affects the natural frequency of the fundamental mode of
computed by taking wind angle of attack into consideration. The mean the building. Previous observation during various levels of earthquake
wind velocity at each story of the building is estimated using the (Siringoringo and Fujino, 2015b) has revealed a relationship between
variation of wind with height formula as follows: the maximum recorded isolator deformation and the identified natural
⎛ z ⎞α frequency of the first mode as shown by a trend line in Fig. 15. The
Ui = Uz 0i ⎜ i ⎟ cos θi figure also shows that the maximum static deformations of isolator
⎝ z0 ⎠ (6)
estimated from wind-induced by MWL and ESWL methods during the
where Uz 0i is the mean of standard wind measured at height z0=10.2 m largest wind speed of the three events. The values are plotted with
(Japan Meteorological Agency (JMA), 2016), and zi is the respective respect to the identified natural frequency of the first mode.
height of each floor, θi is the mean of relative angle between wind mean Note that while the total maximum of isolator displacement due to
direction and the angle parallel to main building's x-axis. The mean of wind-induced vibration is a summation of static and dynamic deforma-
wind angle of attack is assumed constant along the building height. tion, the static component is expected to dominate the total maximum
Furthermore, the static deformation of isolators associated with the displacement, so it is reasonable to compare the maximum recorded
total shear force are estimated assuming the designed lateral stiffness isolator deformation from earthquakes with the maximum static
of isolators Kiso =727,560 kN/m, as: deformation of isolator estimated from wind-induced responses.
The trend line shown in Fig. 15 is obtained from several moderate
VMWL, ESWL earthquakes in 2011 and a large one that is the main shock of March
δisolator =
Kisolator (7) 11, 2011 earthquake where the first frequency dropped to 0.448 Hz
and the maximum recorded isolator displacement is about 5.50 cm.
For the MWL approach, the total shear force is computed by
The more moderate earthquakes cause smaller isolator displacements,
estimating the frontal area (Ai) of each story based on design drawing
generally between 0.15 and 0.30 cm. The results shown in Fig. 15
and taking the value of drag coefficient Cd as 1.08 as suggested for the
reveal that wind-induced maximum isolator deformations estimated
type of building geometry by AIJ-RLB-2004. Meanwhile, for the ESWL
from MWL method are generally in the good agreement with the trend
method, calculation of the gust loading factor GD is conducted using the
line established previously from earthquake records. More specifically,
procedure described in AIJ-RLB-2004. Parameters for gD, Cg, Cg’ and
the results of maximum static deformations of isolator estimated from
RD are all selected using terrain category III and their associated values
wind-induced vibration during the largest wind speed of the three
as defined in the AIJ-RLB-2004 code (AIJ-RLB, 2004). The mode-
events are comparable with the results from moderate earthquakes
shape dependent correction factor ϕD is computed by using the first
shown in Table 4. The results from ESWL, however, are slightly off the
mode identified from system identification for each typhoon event.
trend line. As mentioned previously, the total shear force and the
Note that since the analysis is aimed at finding the maximum wind load
estimated isolator displacement resulting from ESWL method are
and maximum isolator displacement at each typhoon event, the
consistently larger than that of MWL as the consequence of including
frequency and damping ratios associated with the first mode is used
the gust loading factor, whose values are between 2 and 3; while
to calculate the gust load factor since it is the mode that most
natural frequencies remain the same. As the result, the maximum static
significantly affected by isolator deformation (Siringoringo and
deformations of isolator estimated by ESWL method are slightly off the
Fujino, 2015a).
trend line.
Table 4 lists the results of estimates of total shear force and isolator
The results suggest that drag force acting on the building due to
displacement during the period of peak wind speed for the three events
mean wind part induces initial static deformation on the isolators. The
computed by MWL and ESWL methods. The results show that the total

Table 4
Estimated maximum total shear force acting on the isolators resulted from static wind force computed by MWL and ESWL methods, recorded moderate earthquakes and the
corresponding isolator maximum displacements.

Estimated Values Typhoon roke Bomb-cyclone Typhoon guchol

18:30 JST [28.5 m/s] 21:10 JST [22.7 m/s] 00:20 JST [26 m/s]

MWL ESWL MWL ESWL MWL ESWL

Total Shear Force (V) [tonf] 1521.32 4417.58 1363.55 3862.07 1593.86 4679.61
Max Isolator Disp. (cm) 0.209 0.607 0.187 0.531 0.219 0.643
Frequency of 1st mode 0.565 0.565 0.57 0.57 0.576 0.576

Eq. (1) Eq. (2) Eq. (3) Eq. (4) Eq. (5) Eq. (6)
16-4-2011 15-7-2011 25-7-2011 (1) 25-7-2011(2) 21-4-2011 31-7-2011
Max Isolator Disp. (cm) 0.276 0.282 0.176 0.229 0.269 0.2234
Frequency of 1st mode 0.536 0.545 0.572 0.548 0.543 0.550

195
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

0.65

0.6

EQ
0.55
EQ-Trend Line
Roke(MWL)
0.5 Roke(ESWL)
BombCy(MWL)
BombCy(ESWL)
0.45 Guchol(MWL)
Guchol(ESWL)

0.4
0.1 0.2 0.3 0.4 0.5 0.6 0.8 1.0 2.0 3.0 4.0 5.0 6

Fig. 15. Comparison between identified frequency of the first mode and isolator maximum displacement obtained from earthquake measurements (Siringoringo and Fujino, 2015b) and
estimated from wind-induced responses (MWL) and (ESWL) method.

Main Bld Mid Column Main Bld Mid Column Annex Bld X-Axis
Main Bld Outer Column Main Bld Outer Column Annex Bld Y-Axis
14 14 8
EQ1
7 EQ1
12 12
EQ2
6 EQ2
10 10 EQ3
5 EQ3
8 8 EQ4
Floor
Floor

Floor
4 EQ4
6 6
3
4 4
2
2 2 1

0 0 0
-1 -0.5 0 0.5 1 -1 -0.5 0 0.5 1 -1 -0.5 0 0.5 1
(a) Modal Disp Ratio X-Axis (b) Modal Disp Ratio Y-Axis (c) Modal Disp Ratio

14 14 8 Roke
Roke
12 12 7
B.Cyc
B.Cyc
10 10 6
Guchol
5 Guchol
8 8
Floor

Floor

Floor

4
6 6
3
4 4
2
2 2
1
0 0
-1 -0.5 0 0.5 1 -1 -0.5 0 0.5 1 0
-1 -0.5 0 0.5 1
(d) Modal Disp Ratio X-Axis (e) Modal Disp Ratio Y-Axis (f) Modal Disp Ratio

Fig. 16. Comparison of mode shapes of the first mode identified from moderate earthquakes: (a) main building x-axis, (b) main building y-axis and (c) annex building; with the
corresponding mode shapes identified from typhoons and bomb-cyclone: (d) main building x-axis, (e) main building y-axis and (f) annex building (Note: detail on selected earthquakes
and typhoons events are as explained in Table 3).

initial isolator deformation due to static wind force become the original are in the good agreement with the corresponding mode-shapes
point of isolator displacement caused by building vibration associated obtained previously from earthquake records. The results imply that
with the fluctuating wind part. Due to this fact, frequency of wind- not only the natural frequencies and estimated maximum isolator
induced building vibration is not equal to the frequency at the initial displacement but also the identified mode-shapes of obtained from the
stiffness of the isolators but already affected by their initial static wind-induced responses analysis are consistent; suggesting that ap-
deformations. In the case when wind velocity is high for a relatively proximation of the maximum deformation of isolator by computing the
long period of time and wind acting on the same direction as in the case total shear force is satisfactory.
of present analysis, the incremental static deformation of isolators
could accumulate and resulted in significant initial displacement. 7. Conclusions
In addition to natural frequency, mode shapes of the first mode
obtained from the moderate earthquakes and during the largest wind of Analyses of wind-induced responses and dynamics characteristics
the three events are evaluated as illustrated in Fig. 16. One can observe of an asymmetrical base-isolated building during the passage of
clearly that the mode shapes obtained from the wind responses analysis Typhoon Roke 2011, Typhoon Guchol 2012 and a Bomb Cyclone

196
D.M. Siringoringo, Y. Fujino Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 183–197

April 2012 have been presented in this paper. Based on detail analyses Campbell, S., Kwok, K.C.S., Hitchcock, P.A., 2005. Dynamic characteristics and wind-
induced response of two high-rise residential buildings during typhoons. J. Wind
of the measured data, conclusions of the study are summarized as Eng. Ind. Aerodyn. 93 (6), 461–482.
follows: Chen, J., Xu, Y.L., 2004. On modelling of typhoon‐induced non‐stationary wind speed for
tall buildings. Struct. Des. Tall Spec. Build. 13 (2), 145–163.
Chen, Y., Ahmadi, G., 1992. Wind effects on base-isolated structures. J. Eng. Mech. ASCE
1. Wind-induced accelerations of the building increased with the 118 (8), 1708–1727.
increase of approaching wind speed. The relationship between wind Cohen, L., 1995. Time-Frequency Analysis 299. Prentice Hall PTR, Upper Saddle River,
speed and building acceleration can be described by power equa- New Jersey.
Henderson, P., Novak, M., 1989. Response of base‐isolated buildings to wind loading.
tions. Observation from measured wind direction revealed that the Earthq. Eng. Struct. Dyn. 18 (8), 1201–1217.
most significant building responses during the typhoon were the Japan Meteorological Agency (JMA), 2016. Regional Meteorological Observatory List of
along-wind responses. the Automated Meteorological Data Acquisition System (AMeDAS).
Japan Society of Seismic Isolation (JSSI), 2009. Guideline of Wind-resistant Design for
2. Responses of the building were dominated by the first fundamental
Base-isolated Buildings (in Japanese).
flexural mode slightly coupled with torsion. Due to building's Jeary, A.P., Ellis, B.R., 1983. On predicting the response of tall buildings to wind
asymmetric configuration, influence of the second mode, a torsional excitation. J. Wind Eng. Ind. Aerodyn. 13 (1–3), 173–182.
mode, on the overall responses of the floors on the west corner Juang, J.N., 1997. System realization using information matrix. J. Guid. Control Dyn. 20
(3), 492–500.
(inner corner) of the L-shape building is quite significant. Kareem, A., 1997. Modelling of base-isolated buildings with passive dampers under
Contributions of other higher flexural and torsional modes on the winds. J. Wind Eng. Ind. Aerodyn. 72, 323–333.
total responses were found to be insignificant. Kochiyama O., Ikenaga M., Nakamura T., Kaneko S., Takenaka Y., Yoshikawa K., Wada
A., 2004. Study on LRB performance of high-rise base-isolated building during
3. Wind-induced responses of the building revealed non-linear char- strong wind. In: Proceedings 13th World Conference on Earthquake Engineering,
acteristics. Natural frequencies of the fundamental modes (1st and Aug. 2004.
2nd modes) decreased with the increase of amplitude of building Li, Q.S., Yang, K., Wong, C.K., Jeary, A.P., 2003. The effect of amplitude-dependent
damping on wind-induced vibrations of a super tall building. J. Wind Eng. Ind.
acceleration. During the peak wind speed of typhoons and bomb Aerodyn. 91 (9), 1175–1198.
cyclone, the drops of 8–10% of natural frequencies were observed. Li, Q.S., Fang, J.Q., Jeary, A.P., Wong, C.K., Liu, D.K., 2000. Evaluation of wind effects
4. Damping estimates of the fundamental modes increased with the on a supertall building based on full‐scale measurements. Earthq. Eng. Struct. Dyn.
29 (12), 1845–1862.
increase of building accelerations. The increase in damping is
Liang, B., Shishu, X., Jiaxiang, T., 2002. Wind effects on habitability of base-isolated
thought to be contributed by isolators and dissipative energy devices buildings. J. Wind Eng. Ind. Aerodyn. 90 (12), 1951–1958.
that were activated due to wind-induced initial deformation. Magalhães, F., Cunha, Á., Caetano, E., Brincker, R., 2010. Damping estimation using free
decays and ambient vibration tests. Mech. Syst. Signal Process. 24 (5), 1274–1290.
5. Shapes of the fundamental modes identified from wind-induced
Sato D., Suzuki, H., Tamura, T., Fugo,Y., Nakamura, O., Kasai K., Kitamura H., 2012.
vibration had different pattern compared to that of the large earth- Evaluation of Wind-Induced Responses of Isolated Highrise Building Based on
quake, but show similar pattern with the ones from moderate Monitoring Records. In: Proceedings 20th Natl. Symp. on Wind Engineering, Tokyo,
earthquakes. Due to asymmetrical characteristics of the building, pp. 365–370 (in Japanese).
Siringoringo, D.M., Fujino, Y., 2008. System identification applied to long‐span cable‐
small participation of torsional motion on the upper stories, which supported bridges using seismic records. Earthq. Eng. Struct. Dyn. 37 (3), 361–386.
was not observed during large earthquakes, was observed on the first Siringoringo, D.M., Fujino, Y., 2015a. Seismic response analyses of an asymmetric base‐
mode during the wind-induced vibration. Characteristics of the isolated building during the 2011 Great East Japan (Tohoku) Earthquake. Struct.
Control Health Monit. 22 (1), 71–90.
mode-shapes were also found to be strongly influenced by the wind Siringoringo, D.M., Fujino, Y., 2015b. Long‐term seismic monitoring of base‐isolated
speed. Larger modal displacements of the main building's lower building with emphasis on serviceability assessment. Earthq. Eng. Struct. Dyn. 44
stories, which associated with isolator displacement, were observed (4), 637–655.
Takenaka Y., Yoshie K. Kitamura H., Ohkuma T., 2009. Study on wind-resistant design
during the periods of largest wind speed. method of seismically base-isolated buildings. In: Proceedings JSSI 15th
6. The analyses show that reduction in natural frequencies and change Anniversary International Symposium on Seismic Response Controlled Buildings for
in mode shapes during strong wind are related to deformation of Sustainable Society, Tokyo, September 16–18.
Takenaka Y., Suzuki M., Yoshikawa K., Nakamura T., Kouchiyama O., Ikenaga M., 2004.
isolators due to static wind force. The wind-induced maximum Simplified prediction method of wind-induced response of base-isolated tall building
deformations of isolators were estimated by computing the total using LRB. In: Proceedings 18th Natl. Symp. on Wind Engineering, Tokyo, pp. 365–
shear force on the isolators using the mean wind load (MWL) and 370 (in Japanese).
Tamura, Y., Suganuma, S.Y., 1996. Evaluation of amplitude-dependent damping and
the equivalent static wind load (ESWL) approaches. The results were
natural frequency of buildings during strong winds. J. Wind Eng. Ind. Aerodyn. 59
compared with the deformations measured during moderate earth- (2), 115–130.
quakes and better agreements were obtained for the results of the van Overschee, Peter, De Moor, Bart, 1996. Subspace Identification for Linear Systems:
mean wind load (MWL) than the equivalent static wind load (ESWL) Theory, Implementation, Applications. Kluwer Academic Publishers, P.O. Box 17,
3300 Dordrecht, The Netherlands.
approach. Vulcano, A., 1998. Comparative study of the earthquake and wind dynamic responses of
base-isolated buildings. J. Wind Eng. Ind. Aerodyn. 74, 751–764.
References Xu, Y.L., Chen, J., 2004. Characterizing nonstationary wind speed using empirical mode
decomposition. J. Struct. Eng. 130 (6), 912–920.
Yasui H., Ohkuma T., Koga S., Shimomura S., 2002. Full-scale measurement of wind-
AIJ-RLB, 2004. Recommendations on Loads for Building. Architectural Institute of induced response of a base-isolated midrise building. In: Proceedings 17th Natl.
Japan (AIJ), (in Japanese). Symp. on Wind Engineering, Tokyo, pp. 445–450 (in Japanese).
Brincker, Rune, C. Ventura, Palle Andersen, 2001. Damping estimation by frequency
domain decomposition. IMAC XIX, Kissimmee, USA, 9(2001), p. 72.

197

Das könnte Ihnen auch gefallen