Sei sulla pagina 1di 12

Structural Safety 32 (2010) 250–261

Contents lists available at ScienceDirect

Structural Safety
journal homepage: www.elsevier.com/locate/strusafe

Seismic collapse risk of light-frame wood construction considering aleatoric


and epistemic uncertainties
Yue-Jun Yin *, Yue Li
Department of Civil Engineering and Environmental Engineering, Michigan Technological University, 1400 Townsend drive, Houghton, MI 49931, USA

a r t i c l e i n f o a b s t r a c t

Article history: Collapse of light-frame wood buildings in earthquakes causes casualties and economic losses. The col-
Received 26 November 2009 lapse risk of buildings under seismic loads are uncertain because of aleatoric and epistemic uncertainties
Received in revised form 26 February 2010 in both ‘demand’ and ‘capacity’. Both sources of uncertainties are considered in this study to investigate
Accepted 16 March 2010
their effects on the collapse risk of wood structures due to seismic loads. Record-to-record uncertainty
Available online 20 April 2010
and effect of spectral shape (e) of ground motion records are examined. Uncertainties in structural resis-
tance are represented in typical wood-frame shear walls, which are modeled by a hysteresis model with
Keywords:
10 parameters, each of which is treated as a random variable. Epistemic uncertainty that is introduced by
Collapse
Dynamic analysis
the modeling process is examined in this study. The implications of inclusion of all sources of uncertain-
Risk ties on collapse risk are investigated and discussed in the context of comparing with the collapse risk in
Seismic performance concrete and steel structures. It is found that the resistance uncertainty as well as modeling uncertainty
Spectral shape have significant impacts on the seismic collapse risk of light-frame wood buildings. Some previous stud-
Uncertainty ies that neglected the effect of resistance uncertainty in seismic performance evaluation may lead to
Wood construction unconservative results.
Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction curve represents the expected (mean) seismic hazard for a specific
site. Aleatoric uncertainty in earthquake ground motion is inherent
Ninety percent of housing in the United States is light-frame in the randomness of amplitudes, phase angle, and shape of the
wood construction. The damage or collapse of such construction seismic hazard curve, while epistemic uncertainty exists in seismic
under earthquake events caused catastrophic losses. In the 1994 models such as alternate attenuation models [5,6]. In studies using
Northridge earthquake, about $20 billion of property losses [1,2] ground motion records to perform nonlinear dynamic analysis
and 24 out of 25 deaths were due to damage or collapse of wood (NDA), aleatoric uncertainties exist between ground motion re-
construction [1]. In the 1995 Japan Kobe earthquake, collapse of cords [7] and details of each ground motion [8]. In this study,
residential wood buildings contributed significantly to the death uncertainties between ground motion is termed record-to-record
and economic losses [2]. uncertainty, while the ground motion spectral shape parameter e
In order to evaluate the collapse risk of existing buildings or de- [9] is to be investigated for ground motions. Both record-to-record
sign new buildings with certain levels of margin against collapse, it uncertainty and e will be discussed in Section 3. On the capacity
is important to accurately estimate the seismic demand and struc- part, aleatoric uncertainty exists in the damping, stiffness, mass,
tural capacity [3]. However, the inherent uncertainties in earth- and energy dissipation characteristics of the structure, while epi-
quake ground motions as well as structural systems make stemic uncertainty represents how good the numerical model rep-
structural collapse risk evaluation a challenging task. Typically, resents the actual structure as well as how good the drawings
uncertainty due to the inherent randomness is termed aleatoric describe the real structure [10]. In this study, the aleatoric uncer-
uncertainty, while uncertainty due to the limitation of human tainty in structure capacity is termed resistance uncertainty and
knowledge is termed epistemic uncertainty. will be discussed in Section 4. Epistemic uncertainty will be dis-
All sources of uncertainties need to be properly identified, mod- cussed in Section 5.
eled, and propagated in the seismic performance evaluation. A Studies (e.g., [8,11]) indicate that uncertainties of the structural
main source of demand uncertainty is due to seismic loads, which system has limited effect on structural seismic performance. How-
also is the largest source of uncertainty in seismic risk assessment ever, those studies focused mainly on limit states other than col-
[4]. The United States Geological Survey (USGS) seismic hazard lapse risks, which includes minor or moderate damage. Recent
studies [4,6,12,13] indicate that uncertainties associated with
* Corresponding author. Tel.: +1 9063707254. structure capacity (stiffness, strength, energy dissipation, etc.)
E-mail address: yuejuny@mtu.edu (Y.-J. Yin). have significant influence on structural collapse performance of

0167-4730/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.strusafe.2010.03.012
Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261 251

steel and reinforced concrete structures. However, it is unclear main lateral force-resisting system. The configuration of the build-
whether the uncertainty in resistance has a significant impact on ing is shown in Fig. 1, where details of openings for the south shear
the collapse risk of light-frame wood constructions, which is exam- wall are illustrated. The dimensions of the building are 9.75 m
ined in this paper. Some sources of uncertainties were investigated (32 ft) long, 6.10 m (20 ft) wide and 2.44 m (8 ft) high. The shear
for seismic performance of wood construction. For example, Li and walls are covered by 1.22  2.44 m (4  8 ft) sheathing panels,
Ellingwood [1] studied wood shear wall fragilities against three which might be modified, as appropriate, to allow for door and
limit states considering uncertainty from earthquake ground mo- window openings. Studs are spaced at 610 mm (24 in) on centers.
tion. Christovasilis et al. [2] quantified the seismic collapse fragili- For the building located in Los Angeles, the sheathing of the shear
ties of two-story and three-story wood buildings using Increment walls is provided by 11.1 mm (7/16 in) oriented strand board (OSB)
Dynamic Analysis (IDA [14]), where capacity variants were consid- panels. For the other three sites, 9.5 mm (3/8 in) OSB panels are
ered only by examining three object-based levels: poor, typical and used. The sheathing is connected to the studs with 8 d common
superior. Li et al. [15] investigated the collapse probabilities of a nails, which are 3.33 mm (0.131 in) in diameter. For the building
one-story wood building conditioned on the maximum considered located in Los Angeles (building A), 3/6 nailing schedule is used,
earthquakes (MCE, i.e., earthquake with a 2475-year return period) which indicates that nails are spaced 76.2 mm (3 in) along the
and design earthquakes stipulated in ASCE 7-05 [16]. While sheathing panel perimeter and 152.4 mm (6 in) in the panel inte-
considering the uncertainty from earthquake ground motions, rior. For the building in other three sites (building B), 6/12 nailing
these studies have not systematically investigated the effect of schedule is used. The fundamental period of the building A is
aleatoric and epistemic uncertainties for the collapse risk of wood 0.20 s, while it is 0.25 s for the building B.
constructions. The building has a gable roof but not shown Fig. 1a. The roof is
Epsilon (e) was investigated by Baker and Cornell [9] to measure assumed to be rigid. The response of a wood-frame construction
the ground motion intensity and indicate spectral shape of the subjected to seismic load is highly nonlinear and shows pinched
ground motion. Studies on concrete and steel structures by Baker hysteretic behavior with strength and stiffness deterioration.
and Cornell [17], Goulet et al. [18] found that structural collapse Fig. 2 shows the load-displacement response of a typical wood-
capacity determined using ground motion records with larger e frame shear wall subjected to a ground motion record. The back-
tends to be higher than that determined using smaller e ground bone curve (i.e., the envelope of the hysteresis curves) is defined
motion records. The ATC-63 project [19] determined the collapse by:
capacity (defined as a peak inter-story drift of 7%) and the collapse 8
K 0 jdj=F 0
margin ratio for two archetypes (one for residential buildings and < sgnðdÞðF 0 þ r1 K 0 jdjÞð1  e
> Þ if jdj 6 jDu j;
the other for office, retail, warehouse buildings). However, the e ef- F ¼ sgnðdÞF u þ r 2 K 0 ½d  sgnðdÞdu  if jDu j < jdj 6 jDF j; ð1Þ
>
:
fect on structures with a fundamental period less than 0.5 s, which 0 if jdj > jDF j:
is typical for light-frame wood buildings, has not been investigated
[4].
In this study, both aleatoric and epistemic uncertainties from
demand and capacity are propagated into seismic collapse risk 25 (Du,Fu) 1
Hysteresis curve 1 r K
r1K0
assessment of light-frame wood construction. A typical residence 20 Backbone curve 2 0

is assumed to be located at four sites in the United States. Los 15


F
0
Angeles, CA, Seattle, WA, and Boston, MA represent high, moderate
r K
and low seismicity areas, respectively. St. Louis, MO represents 10 3 0
Load, kN

1
sites in the Central and Eastern United States (CEUS) where typi- 5 F
1
1
cally is referred to as low-to-moderate seismicity region [20]. 0
r4K0

The results obtained herein will provide insights into light-frame


wood building collapse risk assessment considering aleatoric and −5
epistemic uncertainties, and provide risk-informed tools for deci- −10
sion making including structural rehabilitation, retrofit or repair −15
K
p
K0
plans. 1
−20
1
−25
2. Structural model −100 −50 0 50 100
Displacement, mm
The collapse risk evaluation is performed on a typical one-story
residential house in the United States with wood shear walls as its Fig. 2. Hysteresis model of light-frame wood shear wall.

(a) 3-D model for the light-frame wood (b) South shear wall (unit: meter)
building (roof is not illustrated)
Fig. 1. Schematic of one-story wood frame residence.
252 Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261

which was developed by Folz and Filiatrault [21]. In Eq. (1), K0 is the (a) IDA drift curves
initial stiffness; Du and Fu correspond to the shear wall ultimate
capacity, after which its load bearing capacity decreases with a 6
slope of r2. The hysteresis curves in Fig. 2 are obtained using the

Spectral acceleration, g
model developed by Folz and Filiatrault [21], who included the 5
backbone function, Eq. (1), and introduced five more parameters
r3, r4, Kp, a, and b. r3 is the unloading stiffness from the backbone 4
curve. r4 is the pinching line slope. Kp is a reloading stiffness consid-
ering stiffness degradation. a and b are two parameters considering 3
stiffness and strength degradation. Details about these parameters
are provided in Folz and Filiatrault [21]. Using this hysteresis model, 2
Folz and Filiatrault [22] developed the SAWS program to perform
NDA for light-frame wood construction. In this study, the hysteresis 1
parameters of the four shear walls in the building are determined
using the CASHEW program [21,23]. The CUREE-Caltech loading 0
protocol [24] is used in the CASHEW program to determine the hys- 0 1 2 3 4 5 6 7
teresis parameters. By comparing the CUREE-Caltech loading proto- Drift, %
col with other protocols, Gatto and Uang [25] concluded that the
CUREE-Caltech loading protocol is appropriate for light-frame wood (b) Collapse fragility without considering
buildings. The dynamic response of the one-story light-frame wood 1
Fitted collapse fragility
building shown in Fig. 1 is obtained using the SAWS program. 0.9 Rank order CDF
Two to four-story light-frame wood buildings are more suscep-
tible to collapse risk, particularly those with garages and large 0.8
openings, than one-story buildings. The one-story building exam- 0.7
ined herein is to demonstrate the relative contribution of various
Probability 0.6
uncertainties to the overall collapse risk. Nevertheless, the numer-
ical model can also be used to model wood building of more than 0.5
one story. This is beyond the scope of the current paper and will be 0.4
investigated in a future study.
0.3
0.2
3. Collapse risk with uncertainty in ground motion records 0.1
0
3.1. Record-to-record uncertainty 0 1 2 3 4 5 6 7 8
Spectral acceleration, g
One of the major challenges associated with performance-based
seismic engineering [26] is the identification, characterization and Fig. 3. IDA and fragility curves of the one-story wood building (building A)
appropriate treatment of the uncertainties in the ground motion subjected to Los Angeles 2/50 ground motions.
prediction [5]. Earthquake events are typically predicted using
attenuation equations (e.g., the one proposed by Abrahamson for building A subjected to a suite of 20 ground motions, which
and Silva [27]), which consider the earthquake magnitude distribu- were developed in the SAC project [28] and has a 2% exceedance
tion, epicenter distance, soil type, rupture type, and other factors. probability in 50 years for Los Angeles. This suite of ground mo-
In order to properly incorporate the inherent uncertainty of earth- tions will be referred to as LA2/50. Similarly, SE2/50, BO2/50, and
quake events in the structural collapse fragility evaluation, IDA has SL2/50 indicate ground motions with 2% exceedance probability
been found to be an effective tool [4,7]. A set of appropriately in 50 years for Seattle, Boston, and St. Louis, respectively. Note that
selected ground motion records (e.g., suites of ground motions the SAC ground motions [28] are used in the calculation for three
developed by the SAC project [28]) are used in IDA and the sites: Los Angeles, Seattle, and Boston, while Wen-Wu ground mo-
record-to-record variability is implicitly accounted for. tion [29] is used for St. Louis.
To perform an IDA, a ground motion record is scaled to multiple The collapse fragility can be defined by a lognormal distribution
levels of intensity with respect to the Sa at the fundamental period [1] as shown by its cumulative distribution function (CDF):
of the structure. NDA is then performed on the structure using each " #
scaled ground motion record. IDA provides a family of IDA curves lnSa  lnSa;cc
F R ðSa Þ ¼ U ð2Þ
describing the relation between intensity measurement (IM, Sa rlnSa;cc
hereby) and engineer demand parameter (EDP, maximum incipient
drift hereby). The point of incipient collapse is reached when a in which the median collapse capacity Sa;cc and the standard devia-
small increment in IM results in a large increment in EDP or the tion of logarithmic collapse capacity rlnSa;cc (also called dispersion of
drift becomes so large (over 7% [2] for wood structure) that the collapse capacity due to record-to-record uncertainty br2r) can be
NDA fails to converge. determined by regression analysis of a set of Sa,cc determined by
Analysis shows that the drift of the south and north shear walls IDA. For Los Angeles, 20 collapse capacities Sa,cc were determined
are larger than those of the west and east shear walls. Ground mo- from Fig. 3a. It is found that the lognormal distribution is an appro-
tions are only applied along the east-west direction to estimate the priate model for the collapse capacity, Sa,cc, of the light-frame wood
building’s collapse risk. For a given ground motion, the observed building using Kolmogorov-Smirnov test at 5% significance level.
(based on dynamic analysis results) incipient drifts are very close Fig. 3b shows the collapse fragility obtained for the south shear wall
for south and north shear walls. In the following sections, all the of the residence in Los Angeles. The dispersion of the collapse
collapse capacity/fragilities are determined using the drift for the capacity is found to be 0.39, which indicates the effect of the re-
south shear wall. Fig. 3a shows 20 IDA curves for drifts obtained cord-to-record uncertainty of the LA2/50 ground motions.
Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261 253

3.2. Effects of spectra shape of ground motion records


a 3.5
on collapse capacity Median attenuation prediction
Median +/− sigma
3 Response spectra

Spectral acceleration, g
Other than the typical IM such as Sa discussed in Section 3.1, re- 2.5
cent studies investigate which properties of a ground motion are
most strongly related to the structural response. Baker and Cornell 2
[9] found that the epsilon parameter (e) is an important parameter 1.5
to measure the intensity of a ground motion. Epsilon indicates the
spectral shape of a ground motion and is a key characteristic of 1
ground motions affecting collapse risk assessment [17]. Recent 0.5
studies [17,18,30,31] have found that selecting ground motions
with peaked spectral shapes increases the predicted structural col- 0
lapse capacity. However, these studies focused on steel or concrete −0.5
structures with fundamental period around 1.0 s. The e effect on
light-frame wood residences, typically with period between 0.2 −1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
and 0.5 s, has never been investigated.
For a given ground motion with a spectral acceleration Sa(T) at Period, sec
period T, the corresponding epsilon can be calculated by:
b 4.5
Subset1
lnðSa ðTÞÞ  E½ln Sba ðTÞ 4 Subset2
eðTÞ ¼ ð3Þ Expected
Std½ln Sba ðTÞ

Spectral acceleration, g
Others of LA2/50
3.5

where E½ln Sba ðTÞ and Std½ln Sba ðTÞ are the mean and standard devia- 3
tion of logarithmic spectral acceleration, ln Sba ðTÞ, predicted by an
2.5
attenuation equation. For a given ground motion record, e is a func-
tion of period and attenuation equations [9]. The attenuation equa- 2
tion proposed by Abrahamson and Silva [27] is used in this study.
1.5
Using a different attenuation model for calculating e might only
lead to a slightly different value [17]. 1
Fig. 4a shows the acceleration spectra of a ground motion re-
0.5
cord with a positive e at period T = 1.0 s. The e value of a ground
motion record at a period T indicates the distance between its 0
spectra and the expected spectra predicted by attenuation models. 0.01 0.1 0.2 0.5 1.0
For example, in this case, e(1.0) > 0 indicates that this ground mo- Period, sec
tion record has a larger spectral acceleration than the expected
Fig. 4. (a) Acceleration spectra of a ground motion with a positive e and (b) spectra
spectral acceleration. Such a ground motion record is termed ‘Peak’
of LA2/50 ground motions and expected spectra using the attenuation model by
record [9]. It has been found that the spectra of one ground motion Abrahamson and Silva [27].
record with a positive e value at a period tends to be peak shaped
around that period [17].
A set of randomly selected ground motions will have an average forced concrete structures. Note that only three ground motions
e value around zero and a unit standard deviation [9]. However, are used in each subset. Accurate quantification of the effect of e
this does not hold for suites of ground motions such as LA2/50, on structural collapse capacity requires more ground motions with
which are intentionally selected so that they represent a certain specific e at the fundamental period of the building of interest. In
seismicity level for a given site (e.g., 2% exceedance probability in Fig. 5, there is one fragility curve that is adjusted with a target e va-
50 years for Los Angeles). Using Eq. (3), the e values for the LA2/ lue of 2. The target e is discussed in Section 3.3 in which the col-
50 ground motions are determined and listed in Table 1. The mean lapse fragility adjustment technique is introduced.
e value at 0.20 s (i.e., the fundamental period of building A in Los The e values (listed in Table 1) for LA2/50 ground motions have
Angeles) of LA2/50 is 0.38 instead of zero for the reason discussed a median value of 0.69 (at the period of 0.2 s), while the median va-
above. Fig. 4b shows the spectra of LA2/50 ground motions, in lue is 1.38 if e is calculated at period of 0.25 s. The reason of the dif-
which the expected spectra is determined by the attenuation mod- ference is that e is a function of period [9]. There is quite a
el by Abrahamson and Silva [27]. Subsets 1 and 2 in Fig. 4b are se- dispersion of e among the ground motions. The reason of the dis-
lected in order to compare collapse fragilities determined from persion may be due to the fact that the SAC ground motions [28]
ground motions with different e. Subset one consists of LA40, were developed by matching the uniform hazard spectra for each
LA24, LA39, and LA38 and has an e range of 1.71 to 1.15. Subset site only in terms of Sa, without considering the spectra shape of
two includes LA28, LA30, LA34, and LA25 and has an e range of each selected ground motion.
1.06–1.39. For each subset, collapse fragility can be obtained using
the method discussed in Section 3.1. Collapse fragilities deter- 3.3. Collapse capacity with adjustment for epsilon
mined using subsets one and two are shown in Fig. 5. The median
collapse capacity obtained using subset two is 2.94 g, 16% larger In Section 3.2, it has been found that e has a significant effect on
than 2.52 g determined without considering e. The median collapse the collapse capacity evaluation of light-frame wood construction.
capacity obtained using subset one is 2.15 g, 14% smaller than It is necessary to properly account for the +e effect when evaluating
2.52 g. From the comparison, it can be observed that larger e can structural collapse risk. One approach is to select a set of ground
lead to a larger collapse capacity prediction (i.e., lower probability motion records with a specific e value at a specified period (e.g.,
of collapse) while smaller e leads to smaller collapse capacity pre- the fundamental period), and then apply IDA methods to deter-
diction (i.e., higher probability of collapse). This is consistent with mine Sa,cc. However, this approach is not feasible for practical use
what has been found in studies [9,17,18,30,31] on steel and rein- because it might be difficult to obtain enough ground motion re-
254 Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261

Table 1
Epsilon (e) values of LA2/50 ground motions (at the period of 0.20 s).

Ground motion e Ground motion e Ground motion e Ground motion e


LA23 2.33 LA37 8.36 e1 LA26 8.99 e1 LA33 1.45
LA40a 1.71 LA22 5.77 e1 LA28 1.06b LA36 1.54
LA24a 1.36 LA21 1.12 e1 LA30 1.10b LA35 2.10
LA39a 1.34 LA27 1.44 e3 LA34 1.22b LA32 2.81
LA38a 1.15 LA29 4.86 e1 LA25 1.39b LA31 3.00
a
Subset one.
b
Subset two.

1 2
0.9 1.8
0.8 1.6
0.7
1.4
Probability

0.6

a,cc
1.2
y=0.12*x+0.88
0.5

lnS
1
0.4
0.8
0.3
0.6
0.2 Without epsilon
−1.71 < epsilon < −1.15
0.1 1.06 < epsilon < 1.39 0.4
Adjusted with epsilon = 2
0 0.2
0 1 2 3 4 5 6 7 8 −3 −2 −1 0 1 2 3
Spectral acceleration, g Epsilon

Fig. 5. Collapse capacity of building A subjected to LA2/50 ground motions (target Fig. 6. Regression of e and lnSa;cc for LA2/50 ground motions (at the period 0.2 s).
e = 2).
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r^ lnðSa;cc Þ ¼ r2lnððSa;cc Þ;regÞ þ c21 r2e ð6Þ
cords with given target e values at specific periods [32]. An alterna-
tive approach was proposed by Haselton et al. [32], which uses a
where c1 is the regressed coefficient in Eq. (4), re the standard devi-
general set of ground motions without considering e and then ad-
ation of e, and rlnððSa;cc Þ;regÞ is the standard deviation of the regression
justs the results distribution to account for e in the analysis of col-
residual of lnðSa;cc Þ.
lapse risk.
For example, for the LA2/50 ground motion records, c0 = 0.88
In the beginning of the second approach, one needs to select a
and c1 = 0.12, the standard deviation re ¼ 1:52, and rlnððSa;cc Þ;regÞ ¼
general set of ground motions without considering e and deter-
0:34. If etarget = 2.0, the adjusted lognormal parameters for collapse
mine the e value and collapse capacity Sa,cc for each ground motion
fragility can be calculated as
record. Then regression analysis between Sa,cc and e is performed
according to: l^ lnðSa;cc Þ ¼ 0:88 þ 0:12  2:0 ¼ 1:12 ð7Þ
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
lnSa;cc ¼ c1 e þ c0 ð4Þ
r^ lnðSa;cc Þ ¼ 0:342 þ 0:122  1:522 ¼ 0:39 ð8Þ
where c1 and c0 are parameters determined by regression analysis.
The two parameters l ^ lnðSa;cc Þ ¼ 1:12 and r
^ lnðSa;cc Þ ¼ 0:39 define the
Fig. 6 shows the regression results for LA2/50 ground motions. Coef-
adjusted collapse fragility curve shown in Fig. 5. The expected
ficients c0 and c1 in Eq. (4) is found to be 0.88 and 0.12, respectively.
median collapse capacity with e = 2 is 3.06 g, 21% larger than the
The coefficient of determination, R2, is found to be 0.24. In this
median collapse capacity 2.52 g without considering e. By perform-
study, c1 and c0 will be taken as deterministic values for a specific
ing the same process for Seattle, Boston, and St. Louis, collapse fra-
site. Note the e of LA2/50 is calculated at the period of 0.20 s (i.e.,
gilities of the residence with considering e are obtained and listed
the fundamental period of building A. e for the other three sites
in Table 2. By comparing the collapse fragility curves in Fig. 5, it can
should be calculated at the period of 0.25 s (i.e., the fundamental
be observed that using this alternative approach to account for e
period of building B. Let fSa;cc g denote
n o the observed collapse capac- effect in collapse fragility is effective. This alternative method is
ity vector for the building and S0a;cc denote the calculated vector
used later in this study to assess the collapse risk of the light-frame
by Eq. (4) using the envalues o listed in Table 1. A residual vector is wood building.
obtained by fSa;cc g  S0a;cc and will be used for collapse fragility
The dispersion parameter rlnSa;cc listed in columns (4) and (6) of
adjustment discussed later in this section. The last step is to adjust
Table 2 reflects the record-to-record uncertainty as discussed in
the collapse fragility using the regression results. Assuming that the
Section 3.1. Since the same suite of ground motions is used to
collapse capacity, Sa,cc, is lognormally distributed, the expected nat-
determine the collapse fragility, the dispersions are the same for
ural logarithm of the collapse capacity, l ^ lnðSa;cc Þ , can be obtained as:
each site regardless whether e is considered or not. Although the
l^ lnðSa;cc Þ ¼ c0 þ c1 etarget ð5Þ MCE spectra acceleration Sa for Seattle and St Louis are 1.57 g
and 0.55 g, respectively, the median e values observed for SE2/50
where c0 and c1 are the parameters shown in Eq. (4), while etarget is and SL2/50 are close (3.00 for SE2/50 and 3.07 for SL2/50). The
the target e of interest. The standard deviation of the natural loga- suites of ground motions (LA2/50, BO2/50, and SE2/50) were devel-
rithm of the collapse capacity, r^ lnðSa;ccÞ , can be determined as: oped to be used for in the SAC project [33]. While the ground
Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261 255

Table 2
Collapse fragilities for the building located at four sites.

Site MCE (g) Median e Coefficients in Eq. (4) Fragility parameters without e Fragility parameters adjusted with e = 2
c0 c1 Sa;cc (g) rlnSa;cc Sa;cc je (g) rlnSa;cc je
(1) (2) (3) (4) (5) (6) (7) (8) (9)
Los Angeles 2.34 0.69a 0.39 0.24 2.52 0.39 3.06 0.39
Seattle 1.57 3.00b 0.47 0.10 2.11 0.41 1.98 0.41
Boston 0.28 1.31b 0.95 0.12 3.19 0.50 3.28 0.50
St. Louis 0.55 3.07b 1.09 0.12 4.31 0.31 3.76 0.31

Sa;cc ; Sa;cc je: median collapse capacity without or with considering e.rlnSa;cc ; rlnSa;cc je: dispersion due to record-to-record uncertainty without or with considering e.
a
at the period of 0.2 s.
b
at the period of 0.25 s.

motions provide satisfactory evaluation for seismic performance displacement Du is reached, as shown in Fig. 2. Second, there is a
(e.g., [1,34]), collapse fragility obtained from these ground motions 95% probability that a normal distribution assigns values in the
might need to be adjusted to account for e effects in order to appro- range of l ± 2 r (l is the mean value and r is the standard devia-
priately estimate the collapse risk. tion). So sampling in the range of l ± 2 r can provide a high enough
confidence level and save computation resources. The bounding
ranges listed in Table 3 are set considering both factors. Since there
4. Collapse capacity considering resistance uncertainty
is no enough experimental data, the COV of each hysteresis param-
eter is assumed to be 0.5 herein. Vamvatsikos and Fragiadakis [7]
Past studies on seismic performance of light-frame wood struc-
examined the resistance uncertainty of a steel frame and assumed
tures (e.g., [1,2,35]) used deterministic hysteresis models, in which
COV of the hysteresis parameters to be 0.4.
the resistance uncertainty has not been considered. In this section,
the resistance uncertainty in light-frame wood buildings is to be
examined. The uncertainty inherent in the structural system have 4.2. Monte Carlo simulation with Latin Hypercube Sampling
been investigated through several approaches including sensitivity
analysis [8,36,37], first-order-second-moment (FOSM) methods As discussed in Section 3.1, IDA requires many runs of NDA. In
and Monte Carlo methods. Sensitivity analysis indicates the this study, 40 runs of NDA are performed for each ground motion.
parameter effects on the structural performance to a limited depth. A single collapse fragility requires 800 runs of NDA using the LA2/
FOSM methods can be used to propagate modeling uncertainties to 50 ground motions (20  40 = 800). A classic Monte Carlo simula-
evaluate the effect on collapse capacity, but might be problematic tion requires a large sampling number N so that the target variable
when the limit state functions are highly nonlinear [38]. Monte (i.e., structural collapse fragility in this section) distribution can be
Carlo method is the most comprehensive but the most computa- estimated. By sampling N times from the parameter distributions
tion expensive approach [7]. listed in Table 3, Monte Carlo simulation generates a population
of N shear walls. Each of the N shear wall needs 800 runs of NDA
4.1. Resistance uncertainty to obtain a collapse fragility. The computational demand can be
overwhelming given a large number N.
The light-frame wood construction has several wood shear The classic Monte Carlo simulation can be further improved by
walls as its main lateral force resistance system. The resistance of the Latin Hypercube Sampling (LHS) technique [39]. The sampling
each shear wall is reflected in 10 hysteresis parameters as dis- number N can be significantly reduced while keeping the same
cussed in Section 2. In Section 3, these hysteresis parameters are accuracy level as a classic Monte Carlo simulation [7]. Generally,
taken as constants (median values) when IDA is performed to ob- LHS randomly selects n different values from each of k variables
tain collapse fragilities. In this section, the uncertainty associated X1, X2, . . ., Xk. The range of each variable is divided into n nonover-
with these hysteresis parameters (referred to as resistance uncer- lapping intervals, each of which has equal area under its probability
tainty later) are investigated. The numerical model of the light- density function (PDF) curve. One value is selected at random from
frame wood building (Fig. 1) includes four shear walls (to be each interval with respect to its PDF. So a n  k matrix of samples
modeled with 40 random variables). In order to alleviate the com- are generated using LHS for the k variables. The ith column contains
putation burden, an individual wood shear wall (the south shear n randomly selected samples for variable Xi. Fig. 7 shows the LHS
wall in the building B) instead of the whole building is investigated process. More details about LHS can be found in Iman et al. [40,41].
for the variation of resistance, which then is used in the whole The hysteresis parameters for the light-frame wood shear wall
building analysis. This simplified procedure is made on the are statistically and/or physically correlated [37]. However, the
assumption that the dispersion of collapse fragility, in terms of available literature does not provide any information on the corre-
coefficient of variance (COV), due to the resistance uncertainty lations. Recent studies on steel and concrete structures (e.g., Vam-
found in the individual shear wall is at the same level as that in vatsikos and Fragiadakis [7], Celik and Ellingwood [13]) assume
the whole building. that each parameter is independent from others. In this study the
As there is no data available for determining the probability dis- same assumption is made and each hysteresis parameter is exam-
tribution of the hysteresis parameters. The normal distribution is ined independently. While one hysteresis parameter is sampled
assumed in this study for each of the hysteresis parameters. Table 3 using LHS, the other nine hysteresis parameters remain unchanged
lists the distribution parameters. The median values of these and hold their mean values. Each hysteresis parameter is sampled
parameters are determined using the CASHEW program [21,23]. 100 times and so 1000 realizations of a single wood shear wall are
The normal distributions have to be truncated considering the fol- created. For Los Angeles, each shear wall realization is analyzed by
lowing factors. First, there are physical limitations on the hystere- IDA using the LA2/50 ground motions. Fig. 8 shows 100 median
sis parameters. For example, parameter r2 should always be less IDA drift curves, each of which corresponds to one realization of
than 0 since it indicates the strength degradation after the ultimate shear wall (modeled by one sample value of r3 from LHS and nine
256 Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261

Table 3
Probabilistic distributions of the hysteresis parameters (Units: kN-mm).

Hysteresis parameter Probabilistic distribution Distribution parameter Truncation range


K0 Truncated normal MK 0 ¼ 2:73; COVK 0 ¼ 0:5 [0.65, 5.46]
Du Truncated normal MDu ¼ 58:9; COVDu ¼ 0:5 [12.0, 117.8]
r1 Truncated normal Mr1 ¼ 2:49; COVr1 ¼ 0:5 [0, 0.166]
r2 Truncated normal Mr2 ¼ 2:54; COVr2 ¼ 0:5 [0.16, 0.02]
r3 Truncated normal Mr3 ¼ 1:29; COVr3 ¼ 0:5 [1, 2.57]
r4 Truncated normal Mr4 ¼ 0:048; COVr4 ¼ 0:5 [0, 0.096]
F0 Truncated normal MF 0 ¼ 24:95; COVF 0 ¼ 0:5 [5.29, 38.24]
F1 Truncated normal MF 1 ¼ 5:29; COVF 1 ¼ 0:5 [0.8, 10.59]
a Truncated normal Ma ¼ 0:73, COVa ¼ 0:5 [0, 1.46]
b Truncated normal Mb ¼ 1:09; COVb ¼ 0:5 [1, 2.18]
M: median value; COV: coefficient of variance

Fig. 7. Latin Hypercube Sampling process.

mean values of other hysteresis parameters). The collapse capacity A tornado chart that indicates the dispersion due to each hys-
Sa;cc of one realization of shear wall is determined using its median teresis parameters is shown in Fig. 9. The vertical axis is located
IDA drift curve that is obtained by the process discussed in Sec- at 1.07 g, which is the median collapse capacity of the shear wall
tion 3.1. For each median IDA curve, a collapse capacity Sa,cc is with all hysteresis parameters holding their mean values. It can
reached when a small increase of Sa leads to a large increase of drift be observed that the effect of each parameter on the collapse
as discussed previously. Each black dot in Fig. 8 corresponds to a capacity of the single wood shear wall varies significantly. The sta-
Sa,cc for a shear wall realization. After 100 collapse capacities Sa,cc tistics of the dispersions of the 10 hysteresis parameters in Fig. 9
are determined, the collapse capacity dispersion due to uncertainty are listed in Table 4, in which the dispersions due to uncertainty
in parameter r3 ; br3 ¼ rlnSa;cc , can then be determined. The same in K0, F0, F1, and a are much larger than that of other parameters.
procedure is performed for the other 9 hysteresis parameters. The first three items (K0, F0, F1) are related to material strength,
Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261 257

1.4 [37]. In this study, viscous damping ratio n = 1% is assumed as con-


stant while the uncertainty in structure system is modeled through
1.2 the hysteresis parameter.
Spectral acceleration, g

Let br indicate the collapse capacity dispersion due to resistance


1 uncertainty that was investigated earlier in this section. br was ob-
tained using the LA2/50 ground motions so that the record-to-re-
0.8 cord uncertainty br2r was included. Using the procedure
discussed in Section 3.1, a single shear wall modeled by 10 hyster-
0.6 esis parameters with their mean values is analyzed and a disper-
sion br2r ¼ 0:30 is obtained, which is due to the record-to-record
0.4 uncertainty. Assuming the 10 hysteresis parameters are indepen-
dent, the collapse capacity dispersion due to resistance uncertainty
0.2
can be determined by Eq. (9) to be 0.18.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0
0 0.5 1 1.5 2 2.5 3 br ¼ b2total  b2r2r ð9Þ
Drift, % where
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Fig. 8. Effect of uncertainty in hysteresis parameter r3 (shown in Fig. 2) on the X10 2
median collapse capacity. btotal ¼ b
i¼1 p;i
ð10Þ

In Eq. (10), bp;i coresponds to the dispersion items due to each of


the 10 hysteresis parameters K0, F0, F1, Du, r1, r2, r3, r4, a, and b (i.e.,
bK 0 ; bF 0 ; bF 1 ; bDu ; br1 ; br2 ; br3 ; br4 ; ba , and bb ). The values of bp;i ; i ¼
1; . . . ; 10 are listed in Table 4.
To determine whether br = 0.18 obtained in Eq. (9) is reason-
able, an alternative method used by Vamvatsikos and Fragiadakis
[7] is also examined here. br can be determined by:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
PN 2
j¼1 ðlnSða;ccÞ;j  lnSa;cc Þ
br ¼ ð11Þ
N1
where N is the number of structure samples (i.e., 1000 realization of
shear walls), Sða;ccÞ;j the median collapse capacity for the jth shear
wall, and lnSa;cc is the mean value of all the ln Sa;cc . br is found to
be 0.17 using Eq. (11). Therefore, dispersion in collapse capacity
due to resistance uncertainty is approximately 0.18.
Fig. 9. Effect of hysteresis parameters on dispersion of collapse capacity.

5. Collapse risk with aleatoric and epistemic uncertainties


while the last item (a) is related to the energy dissipation mecha-
nism of the structure system. It can be observed that the energy 5.1. Collapse fragility with uncertainty
dissipation capacity and material strength have significant influ-
ence on collapse risk assessment of light-frame wood structures. In Sections 3 and 4, aleatoric uncertainties in earthquake re-
In comparison, Vamvatsikos and Fragiadakis [7] found the yield cords and structural resistance are investigated. Another signifi-
moment, the ultimate ductility, the capping ductility, and the neg- cant source of uncertainty is epistemic uncertainty. It comes
ative stiffness ratio have significant impact on steel structure per- from the fact that numerical models (e.g., attenuation equations
formance. For concrete structures, the viscous damping ratio and and hysteresis models in this study) can only capture part of the
concrete strength were found to have the greatest impact on the real system. The epistemic uncertainty is termed modeling uncer-
structural performance evaluation [13]. Celik and Ellingwood tainty in this study. Let bm indicate the dispersion of collapse
[13] investigated the effect of uncertainty in viscous damping ratio capacity due to modeling uncertainty. The square-root-sum-of-
n for reinforced concrete structures. Studies [21,37,42] have found squares (SRSS) method has been used to assess structural perfor-
that the viscous damping ratio of wood frame structures is in the mance in the presence of various sources of uncertainties (e.g.,
range of 0–2%, lower than that of reinforced concrete structures [3,6,7,34,43]). Using SRSS, the dispersion of collapse capacity con-
(e.g., Celik and Ellingwood [13] assumed a mean n value of 4.6% sidering both aleatoric and epistemic uncertainties can be deter-
for reinforced concrete structures). The reason for this is there is mined by:
significant hysteretic damping in wood structures and the hyster- qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
etic damping is typically incorporated into the hysteresis model b¼ b2r2r þ b2r þ b2m ð12Þ

Table 4
Single shear wall collapse capacity Sa,cc statistics due to resistance uncertainty.

Hysteresis parameter Median value (g) rlnSa;cc (%) Hysteresis parameter Median value (g) rlnSa;cc (%)
K0 1.11 15.1 F0 1.11 15.5
Du 1.16 5.3 F1 1.16 14.2
r1 1.15 7.6 r2 1.7 4.1
r3 1.11 8.4 r4 1.15 6.8
a 1.15 16.4 b 1.15 4.4
258 Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261

where b is the collapse capacity dispersion considering both alea- uncertainty and a target e of 2. The probability increases to 35.9%
toric and epistemic uncertainties, br2r the dispersion due to re- while considering record-to-record, resistance, and modeling
cord-to-record uncertainty discussed in Section 3, br is the uncertainties (see column (8) in Table 6). Considering the same
dispersion due to resistance uncertainty discussed in Section 4, cases in Seattle, the collapse probability rises from 28.6% up to
and bm is the epistemic uncertainty. The structural collapse fragility 37.8% (see columns (7) and (8) in Table 6). For St. Louis and Boston,
dispersion parameter rlnSa;cc defined in Eq. (2) is to be updated in the collapse probabilities given an MCE event are almost zero for
this section to account for all sources of uncertainties. Table 5 tab- all the cases, as can be observed from Figs. 10 and 11. Note that
ulates all the related dispersion quantities. In column (9) of Table 5, in this study the same construction details are assumed for the four
the epistemic uncertainty caused dispersion in collapse capacity, sites.
bm, is set with three levels, 0.2, 0.4, and 0.6, which represents small,
moderate, and high modeling uncertainties, respectively. Similar
5.3. Annual and 50-year collapse probabilities
assumptions have been adopted in other studies investigating the
uncertainty effect on seismic performance assessment [10,34]. The
By convolving the collapse fragilities with seismic hazard
annual collapse probability, Pc,1, and the collapse probability in 50
curves, the annual collapse risk can be evaluated. Several studies
years, P c;50 , shown in columns (6)–(7) and (11)–(13) in Table 5, will
([3,6,15,44]) have used Eq. (13) to estimate the annual collapse
be discussed in Section 5.3.
probability. In Eq. (13), Pc,1 is the annual collapse probability, Sa;cc
Fig. 10 illustrates collapse fragilities of the light-frame building
is the median collapse capacity (the spectral acceleration corre-
in Los Angeles, Seattle, Boston, and St. Louis. It can be observed
sponding to 50% probability of collapse fragility curve), b is the dis-
from this figure that all the collapse fragility curves rotate about
persion of the collapse fragility, and k0 and k are two site specific
the median capacities. The reason is that the median capacity for
parameters defining the seismic hazard [3].
each collapse fragility is unchanged. For example, the median col-
lapse capacity for Los Angeles is 2.52 g. The five fragility curves 0:5ðkbÞ 2
Pc;1 ¼ k0 Sk
a;cc e ð13Þ
shown in Fig. 10 for Los Angeles are defined by the median value
50
2.52 g and five dispersions that corresponds to different combina- Pc;50 ¼ 1  ð1  Pc;1 Þ ð14Þ
tions of uncertainties. Similar results are observed in Liel et al. [4],
Using Eqs. (13) and (14), the annual collapse probability and the
ATC [19].
collapse probability in 50 years ðP c;50 Þ of the light-frame wood
building in Los Angeles, Boston, Seattle, and St. Louis are deter-
5.2. Probability of collapse at MCE mined and summarized in Table 5. The median values Sa;cc used
in the calculation are listed in Table 2. Seismic parameter k0 and
The collapse probabilities at MCE of the light-frame wood build- k for the four sites are obtained from the USGS [45]. The collapse
ing at the four sites are summarized in Table 6. In Los Angeles, gi- capacity dispersion, br = 0.18, is assumed to be the same as that
ven an MCE event, the collapse probability of building A in the single shear wall as discussed in Section 4.
considering both aleatoric and epistemic uncertainties is 46.1%, The annual collapse probabilities considering record-to-record
while it is 34.7% for building B in Seattle subjected to an MCE uncertainty and spectral shape effects are listed in column (7) of
event. Note that the median collapse capacity of building A is Table 5, while the annual collapse probabilities considering only
2.52 g, higher than 2.11 g of building B. The difference between record-to-record uncertainty without spectral shape effects are
the MCE collapse probability is due to differences in seismicity be- listed in column (6). With median collapse capacity adjusted with
tween the two sites (e.g., MCE in Los Angeles and Seattle are 2.34 g target e = 2, the annual collapse probability in Los Angeles de-
and 1.57 g, respectively). creases from 0.53e3 to 0.28e3 (or 46% less). For Boston, the dif-
Collapse fragility curves generated considering the e effect are ference between columns (6) and (7) is negligible. This results are
shown in Fig. 11. Given an MCE event (i.e., Sa = 2.34 g) in Los Ange- consistent with what was recommended in other studies [17,32]
les, the collapse probability is 24.6% considering record-to-record that the spectral shape has significant influence on collapse risk

Table 5
Collapse probability for light-frame wood construction at four sites considering aleatoric and epistemic uncertainties.

Site k0 k Sa;cc a br2r Pc,1  103b Pc,1  103c br bm b Pc,1  103d Pc,50  103d Pc,1  103e
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13)
LA 4.66e3 3.20 2.52 0.39 0.53 0.28 0.18 0.2 0.47 0.76 0.38 0.41
0.4 0.59 1.41 0.68 0.76
0.6 0.74 3.93 1.79 2.11
SE 1.2e3 2.15 2.11 0.41 0.36 0.41 0.18 0.2 0.49 0.42 0.21 0.48
0.4 0.60 0.56 0.27 0.64
0.6 0.75 0.88 0.43 1.01
BO 8.4e5 1.46 3.19 0.50 0.02 0.019 0.18 0.2 0.57 0.022 0.011 0.02
0.4 0.67 0.025 0.012 0.024
0.6 0.80 0.031 0.015 0.03
SL 7.44e3 1.66 4.31 0.31 0.76 0.95 0.18 0.2 0.41 0.83 0.41 1.05
0.4 0.54 0.98 0.48 1.23
0.6 0.70 1.29 0.63 1.62

LA: Los Angeles; SE: Seattle; BO: Boston; SL: St. Louis. k0, k: seismic hazard parameters. br2r, br: collapse capacity dispersion considering record-to-record or resistance
uncertainty. bm, b: collapse capacity dispersion considering modeling or all sources of uncertainties. Pc,1: annual collapse probability; Pc,50: collapse probability in 50 years.
a
For convenience, median collapse capacities tabulated in column (7) of Table 2 are listed here again.
b
Only considering record-to-record uncertainty.
c
Only considering record-to-record uncertainty but with target e = 2.
d
Considering both aleatoric and epistemic uncertainties without considering e effect.
e
Considering both aleatoric and epistemic uncertainties with target e = 2.
Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261 259

(a) Los Angeles (b) Seattle


1 0.5

0.8 0.4
Probability

Probability
0.6 0.3

0.4 0.2

0.2 0.1

0 0
0 1 2.34 3 0 0.5 1.0 1.57 2
Spectral acceleration, g Spectral acceleration, g

(c) Boston (d) St Louis


0.35 0.14

0.3 0.12

0.25 0.1
Probability

Probability

0.2 0.08

0.15 0.06

0.1 0.04

0.05 0.02

0 0
0.28 1.0 1.5 2 0 0.55 1.0 1.5 2
Spectral acceleration, g Spectral acceleration, g

Fig. 10. Collapse fragility of the light-frame wood building at four sites (building A in Los Angeles and building B at the other three sites).

Table 6
Collapse probability (%) of the light-frame wood building at four sites in the United 2.5e5 with considering an epistemic dispersion bm = 0.4 and
States at MCE.
resistance uncertainty br = 0.18. The resistance uncertainty and
Source of uncertainty considered epistemic uncertainty do not have significant influence on the col-
br2r lapse risk of light-frame wood construction in Boston. While in Los
br2r br2r br2r br Angeles the annual collapse probability increases 168% from
br2r br br br br2r bm = 0.6 0.53e3 to 1.41e3.
Site br2r br bm = 0.2 bm = 0.4 bm = 0.6 etarget = 2 etarget = 2 Compared with the record-to-record uncertainty, br2r, the resis-
(1) (2) (3) (4) (5) (6) (7) (8)
tance uncertainty contributes to the overall collapse risk, especially
LA 42.5 43.2 43.7 45.0 46.1 24.6 35.9 when the modeling uncertainty is at moderate or high level (i.e.,
SE 23.6 25.6 27.3 31.1 34.7 28.6 37.8
bm = 0.40.6). Note that the dispersion br listed in Table 5 assumes
BO 0 0 0 0 0.1 0 0.1
SL 0 0 0 0 0.2 0 0.3 the resistance uncertainty effect on a single shear wall is the same
as that on a whole wood building. This assumption needs to be fur-
br2r is the collapse capacity dispersion due to record-to-record uncertainty. br is the
ther investigated later. According to these observations, it can be
collapse capacity dispersion due to resistance uncertainty bm is the collapse
capacity dispersion due to modeling uncertainty. seen that the uncertainty from earthquake ‘demand’ and effect of
spectral shape are more significant than the resistance uncertainty,
assessment for high seismicity areas like western United States. For on the collapse risk analysis for light-frame wood structures. How-
low seismicity areas like Boston (MCE = 0.28 g), the spectral shape ever, neglect of resistance uncertainty will lead to unconservative
might not be as important as other factors such as resistance and estimation of seismic collapse risk of the light-frame wood struc-
modeling uncertainties. ture. In comparison, studies on concrete and steel structures (e.g.,
The annual collapse probability in Boston only considering re- [4,6,12,13]) found that uncertainties in structural systems have
cord-to-record was found to be 2e5, which increases by 25% to significant influence on seismic collapse performance.
260 Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261

(a) Los Angeles (b) Seattle


0.9 0.7

0.8
0.6
0.7
0.5
0.6
Probability

Probability
0.5 0.4

0.4 0.3

0.3
0.2
0.2
0.1
0.1

0 0
0 1 2.34 3 0 0.5 1.0 1.57 2
Spectral acceleration, g Spectral acceleration, g

(c) Boston (d) St Louis


0.35 0.2

0.18
0.3
0.16

0.25 0.14
Probability

0.12
Probability

0.2
0.1
0.15
0.08

0.1 0.06

0.04
0.05
0.02

0 0
0.28 1.0 1.5 2 0 0.55 1.0 1.5 2
Spectral acceleration, g Spectral acceleration, g
Fig. 11. Collapse fragility adjusted with e of the light-frame wood construction at four sites (building A in Los Angeles and building B at the other three sites).

6. Discussions and deflection amplification factor Cd) for use in seismic design.
Although the ATC-63 project includes seismic performance evalu-
The record-to-record dispersion br2r found in this study for the ation for light-frame wood buildings, a direct comparison of results
four sites ranges between 0.31 and 0.50. This is comparable to from the ATC-63 project and those from this study is confounded
what has been found by Haselton and Deierlein [46], who evalu- by several factors. First, the focus of this study is different from that
ated the collapse risk of 30 reinforced concrete moment frames of the ATC-63. This paper aims to evaluate effects of uncertainty on
of varying height. In their study, the dispersions due to record- the seismic performance of a typical light-frame wood buildings,
to-record uncertainty, br2r, were found to range between 0.35 while the ATC-63 focuses on providing reliable performance fac-
and 0.45. The modeling uncertainty found in their study [46] was tors for design purpose. The collapse risk of existing buildings is
around 0.45. Note that the modeling uncertainty in their study cor- not considered in the ATC-63 project but is studied in this paper.
responds to the combination of br (column (8) in Table 5) and bm Second, the ground motions used in the two studies are different.
(column (9) in Table 5) in this study using SRSS. Taking a moderate The SAC and Wen-Wu ground motions are used in this study while
epistemic uncertainty, bm = 0.40, a combined dispersion brþm ¼0:44 the ATC-63 used ground motions from the PEER NGA database
is obtained, which is comparable to the results in Haselton and [47]. Different sources of ground motions lead to different collapse
Deierlein [46]. capacities, as can be seen from Table 2. Third, the definitions of col-
The ongoing ATC-63 project [19] is developing a methodology lapse capacity are different. The median collapse capacity in the
for quantifying structural seismic performance parameters (i.e., re- ATC-63 project was obtained by scaling all the records to the
sponse modification coefficient R, system overstrength factor X0 MCE intensity and then increasing intensity until half of the scaled
Y.-J. Yin, Y. Li / Structural Safety 32 (2010) 250–261 261

ground motion records cause collapse. The median collapse capac- [17] Baker JW, Cornell CA. Spectral shape, epsilon and record selection. Earthquake
Eng Struct Dyn 2006;35:1077–95.
ity determined in this study is discussed in Section 3.1.
[18] Goulet CA, Haselton CB, Mitrani-Reiser J, Beck JL, Deierlein GG, Porter KA, et al.
Evaluation of the seismic performance of a code-conforming reinforced-
7. Conclusions concrete frame building – from seismic hazard to collapse safety and economic
losses. Earthquake Eng Struct Dyn 2007;36(13):1973–97.
[19] ATC, Quantification of Building Seismic Performance Factors, ATC-63 project
Collapse risk of light-frame wood construction at four sites in report (90% draft); 2008.
the United States were assessed in this study, considering both ale- [20] Rosowsky DV, Pang W, Wang Y, Ellingwood BR. Fragility of conventional
woodframe structures built in regions of low-to-moderate seismicity. In: 10th
atoric uncertainty (i.e., record-to-record uncertainty, and resis- International conference on structural safety and reliability; 2009.
tance uncertainty) and epistemic uncertainty (i.e., modeling [21] Folz B, Filiatrault A. Cyclic analysis of wood shear walls. J Struct Eng
uncertainty). The spectral shape (e) effect on the collapse risk 2001;127(4):433–41.
[22] Folz B, Filiatrault A. Seismic analysis of woodframe structures. II: model
was investigated. Collapse probabilities at MCE and collapse prob-
implementation and verification. J Struct Eng 2004;130(9):1361–70.
abilities (annual and 50-year) for four sites were estimated. [23] Folz B, Filiatrault A. CASHEW – version 1.0: A computer program for cyclic
Record-to-record uncertainty found in light-frame wood con- analysis of wood shear walls, Tech. Rep., Division of Structural Engineering,
University of California, San Diego, Report No.: SSRP–2000/10; 2000.
struction in this study ranged between 0.31 and 0.50, which is con-
[24] Krawinkler H, Parisi F, Ibarra L, Ayoub A, Medina R. Development of a testing
sistent with what has been found in steel and concrete structures. protocol for wood frame structures, Tech. Rep., Stanford University, Stanford,
The spectral shape e of ground motion was found to have signifi- Calif, CA, CUREE-Caltech Woodframe Project Report No. W-02; 2001.
cant effect on the collapse risk of light-frame wood construction, [25] Gatto K, Uang C-M. Effects of loading protocol on the cyclic response of
woodframe shearwalls. J Struct Eng 2003;129(10):1384–93.
especially for high seismicity areas like the west coast of United [26] Bertero RD, Bertero VV. Performance-based seismic engineering: the need for a
States. Considering a moderate modeling uncertainty (i.e., reliable conceptual comprehensive approach. Earthquake Eng Struct Dyn
bm = 0.4 in this paper), the dispersion due to both resistance and 2002;31(3):627–52.
[27] Abrahamson NA, Silva WJ. Empirical response spectral attenuation relations
modeling uncertainties was found to be approximately 0.44, which for shallow crustal earthquake. Seismol Res Lett 1997;68(1):94–127.
led to an increase of annual collapse probability ranging between [28] SAC steel project, Suites of earthquake ground motions for analysis of steel
25% and 168% depending on the site. Therefore, resistance and moment frame structures; 2000. <http://nisee.berkeley.edu/data/
strong_motion/sacsteel/ground_motions.html>.
modeling uncertainty contributions cannot be neglected in light- [29] Wen YK, Wu CL. Uniform hazard ground motions for mid-America cities.
frame wood construction collapse risk assessment. Earthquake Spectra 2001;17(2):359–84.
[30] Haselton C, Baker JW. Ground motion intensity measures for collapse capacity
prediction: choice of optimal spectral period and effect of spectral shape. In:
Acknowledgement
Proceedings of 8th national conference on earthquake engineering, San
Francisco, California; 2006.
The authors thank Professor William M. Bulleit for his careful [31] Zareian F. Simplified performance-based earthquake engineering, Ph.D. thesis,
Department of Civil and Environmental Engineering, Stanford University;
review and constructive comments on this paper.
2006.
[32] Haselton C, Baker JW, Liel AB, Deierlein GG. Accounting for ground motion
References spectral shape characteristics in structural collapse assessment through an
adjustment for epsilon. J Struct Eng 2010. doi:10.1061/(ASCE)ST.1943-
[1] Li Y, Ellingwood BR. Reliability of woodframe residential construction 541X.0000103.
subjected to earthquakes. Struct Safety 2007;29(4):294–307. [33] Somerville P, Smith N, Puntamurthula S, Sun J. Development of ground motion
[2] Christovasilis IP, Filiatrault A, Constantinou MC, Wanitkorkul A. Incremental time histories for phase 2 of the FEMA/SAC steel project, Tech. Rep., SAC Joint
dynamic analysis of woodframe buildings. Earthquake Eng Struct Dyn Verture; 1997.
2009;38(4):477–96. [34] Yun S-Y, Hamburger RO, Cornell CA, Foutch DA. Seismic performance
[3] Cornell CA, Jalayer F, Hamburger RO, Foutch DA. Probabilistic basis for 2000 evaluation for steel moment frames. J Struct Eng 2002;128:534–45.
SAC Federal Emergency Management Agency steel moment frame guidelines. J [35] Yin Y-J, Li Y. Probabilistic analysis of light-frame wood building subjected to
Struct Eng 2002;128(4):526–33. combined snow and earthquake loads. In: Inaugural international conference
[4] Liel AB, Haselton CB, Deierlein GG, Baker JW. Incorporating modeling of the engineering mechanics institute, Minneapolis, MN, USA; 2008.
uncertainties in the assessment of seismic collapse risk of buildings. Struct [36] Esteva L, Ruiz SE. Seismic failure rates of multistory frames. J Struct Eng
Safety 2009;31(2):197–211. 1989;115(2):268–84.
[5] Bommer JJ, Crowley H. The influence of ground-motion variability in [37] Rosowsky DV. Reliability-based seismic design of wood shear walls. J Struct
earthquake loss modelling. Bull Earthquake Eng 2006;4(3):231–48. Eng 2002;128(11):1439–53.
[6] Ellingwood BR, Kinali K. Quantifying and communicating uncertainty in [38] Baker JW, Cornell CA. Uncertainty propagation in probabilistic seismic loss
seismic risk assessment. Struct Safety 2009;31(2):179–87. estimation. Struct Safety 2008;30(3):236–52.
[7] Vamvatsikos D, Fragiadakis M. Incremental dynamic analysis for estimating [39] McKay MD, Conover WJ, Beckman RJ. A comparison of three methods fro
seismic performance sensitivity and uncertainty. Earthquake Eng Struct Dyn selecting values of input variables in the analysis of output from a computer
2010;39:141–63. code. Technometrics 1979;221:239–45.
[8] Porter KA, Beck JL, Shaikhutdinovc RV. Sensitivity of building loss estimates to [40] Iman RL, Helton JC, Campbell JE. An approach to sensitivity analysis of
major uncertain variables. Earthquake Spectra 2002;18(4):719–43. computer models: part i: introduction, input variable selection and
[9] Baker JW, Cornell CA. A vector-valued ground motion intensity measure preliminary variable assessment. J Qual Technol 1981;13:174–83.
consisting of spectral acceleration and epsilon. Earthquake Eng Struct Dyn [41] Iman RL, Helton JC, Campbell JE. An approach to sensitivity analysis of
2005;34(10):1193–217. computer models: part ii: ranking of input variables, response surface
[10] ATC, Guidelines for seismic performance assessment of buildings, ATC-58 50% validation, distribution effect and technique synopsis. J Qual Technol
draft report; 2009. <http://www.atcouncil.org/pdfs/ATC-58-50percentDr%aft. 1981;13:232–40.
pdf>. [42] Foliente GC. Hysteresis modeling of wood joints and structural systems. J
[11] Lee T-H, Mosalam KM. Seismic demand sensitivity of reinforced concrete Struct Eng 1995;121:1013–22.
shear-wall building using FOSM method. Earthquake Eng Struct Dyn [43] ATC, ATC-58 Project; 2001. <http://www.atcouncil.org/atc-58.shtml>.
2005;117(14):32–40. [44] Luco N, Ellingwood BR, Hamburger RO, Hooper JD, Kimball JF, Kircher CA, Risk-
[12] Ibarra L. Global collapse of frame structures under seismic excitations, Ph.D. targeted versus current seismic design maps for the conterminous United
thesis, Department of Civil and Environmental Engineering, Stanford States. In: SEAOC 76th annual convention. Structural Engineers Association of
University; 2003. California; 2007.
[13] Celik OC, Ellingwood BR. Seismic fragilities for non-ductile reinforced concrete [45] USGS, Earthquake ground motion parameters version: 5.0.9 – 10/06/2008, US
frames – role of aleatoric and epistemic uncertainties. Struct Safety Geological Survey; 2008. <http://earthquake.usgs.gov/research/hazmaps/
2010;32(1):1–12. design/>.
[14] Vamvatsikos D, Cornell CA. Incremental dynamic analysis. Earthquake Eng [46] Haselton CB, Deierlein GG. Assessing seismic collapse safety of modern
Struct Dyn 2001;31(3):491–514. reinforced concrete moment frame buildings, Tech. Rep., Department of Civil
[15] Li Y, Yin YJ, Ellingwood BR, Bulleit WM. Uniform hazard vs. uniform risk bases and Environmental Engineering, Stanford University; 2007.
for performance-based earthquake engineering of light-frame wood [47] Pacific Earthquake Engineering Research Center (PEER), Pacific Earthquake
construction. Earthquake Eng Struct Dyn 2010. doi:10.1002/eqe.989. Engineering Research Center: NGA Database; 2005. <http://peer.berkeley.edu/
[16] American Society of Civil Engineers, Minimum design loads for buildings and nga/>.
other structures (ASCE 7-05); 2006.

Potrebbero piacerti anche