You are on page 1of 6

EARTHQUAKE ENGINEERING & STRUCTURAL DYNAMICS

Earthquake Engng Struct. Dyn. 2015; 44:1157–1162


Published online 22 October 2014 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/eqe.2513

SHORT COMMUNICATION

Record selection for aftershock incremental dynamic analysis

Katsuichiro Goda*,†
Department of Civil Engineering, University of Bristol, Bristol, BS8 1TR, UK

SUMMARY
The back-to-back application of mainshock records as aftershock is often considered in conducting after-
shock incremental dynamic analysis. In such an approach, the characteristics of mainshock records are con-
sidered to be similar to those of major aftershock records within the same mainshock–aftershock sequences.
The underlying assumption is that the characteristics of selected mainshocks, other than those used for record
selection, are not significant in the assessment of structural responses. A case study is set up to investigate the
effects of aftershock record selection on the collapse vulnerability assessment. The numerical results for a
specific wood-frame structure indicate that the aftershock fragility can be affected by the aftershock record
characteristics, particularly response spectral shape. Copyright © 2014 John Wiley & Sons, Ltd.

Received 16 June 2014; Revised 2 October 2014; Accepted 6 October 2014

KEY WORDS: record selection; incremental dynamic analysis; aftershock; back-to-back analysis

1. INTRODUCTION

Seismic risk assessment considering both mainshocks and aftershocks is a timely research topic [1–9].
Numerous studies have been conducted by adopting realistic finite-element models [2, 6–9] as well as
inelastic single-degree-of-freedom systems [1, 3–5]. To represent ground motion records for mainshock–
aftershock (MS-AS) sequences, as-recorded data [4–8] as well as artificial time histories [1–5, 9] have
been used. When MS-AS sequences are synthesized, aftershock records are often substituted by scaled
mainshock records. This is referred to as back-to-back analysis. Alternatively, seismological models,
such as modified Omori law and Gutenberg–Richter law, can be utilized [4, 5]. Moreover, inelastic
seismic demand estimation methods adopted in past studies differ; incremental dynamic analysis (IDA [10])
has been applied using back-to-back analysis for evaluating aftershock fragility and residual capacity
for mainshock-damaged structures [1, 3, 9] and using real MS-AS sequences for evaluating the
combined MS-AS effects on seismic vulnerability [4, 5, 8]. A cloud-type analysis method has been
extended to develop adaptive aftershock vulnerability models that are updated with observed ground
motions successively during an MS-AS sequence [7].
This article is concerned with selection of aftershock ground motion records that are employed in
assessing seismic performance and collapse vulnerability of buildings and infrastructure. In particular,
back-to-back application of scaled mainshock records as aftershock in aftershock IDA [1] is focused
upon. This consideration is motivated by the facts that several extensive numerical studies have been
conducted recently [3, 9], taking advantage of improved access to high-performance computing
resources, and that this trend will continue. Key issues that are discussed in this study include the
following: how aftershock ground motions are defined and selected and how such records, when scaled

*Correspondence to: Katsuichiro Goda, Department of Civil Engineering, University of Bristol, Bristol, BS8 1TR UK.

E-mail: Katsu.Goda@bristol.ac.uk

Copyright © 2014 John Wiley & Sons, Ltd.


1158 K. GODA

up, affect the evaluation of nonlinear response potential due to MS-AS sequences. It is important to
recognize the findings from past studies [11, 12] that investigated the effects of record scaling on
inelastic seismic demand estimation for mainshocks. These studies indicated that the similarity of
response spectral shapes of selected records to the target response spectra is a critical factor. Along the
same line of reasons, it will be demonstrated that generally, record characteristics of mainshocks and
aftershocks within the same sequences are different (thus, their response spectral shapes are different),
and that the response spectral shape is important for aftershock IDA. It is noteworthy that the issue
discussed in this study is focused upon cases where record scaling is implemented on the basis of
spectral acceleration at the fundamental vibration period of an intact structure Sa(T1) (i.e., intensity
measure IM = Sa(T1)). When advanced IMs [13, 14] are adopted, the estimation bias induced by the
response spectral effects is likely to be reduced. More broadly, the problem discussed herein is related
to the sufficiency of IMs in conducting aftershock IDA, which needs further investigations.
A case study, focusing upon a wood-frame house that is typical in southwestern British Columbia,
Canada, is presented to illustrate the effects of aftershock record selection on the collapse vulnerability
assessment. It is not intended as a rigorous assessment for aftershock fragility of this type of
construction, and the obtained results are specific to this case study. Nonetheless, the conclusion
drawn from the case study regarding the impact of aftershock record selection on the seismic
performance evaluation is considered to be qualitatively applicable to other situations and is useful
to the earthquake engineering community.

2. AFTERSHOCK INCREMENTAL DYNAMIC ANALYSIS

2.1. Structural model


A typical single-family, two-story residential wood-frame house in southwestern British Columbia is
used in this study. It is based on a University of British Columbia-Seismic Analysis of Wood-frame
Structures (UBC-SAWS) model [15], model parameters of which were calibrated through extensive
static and dynamic test results of shear walls and full-scale houses [16]. The SAWS is a so-called
pancake model that takes bidirectional horizontal seismic excitation into account but not vertical
excitation or P-Δ effects. Building diaphragms are considered to be rigid, and nonlinear hysteretic
behavior of a shear-wall element is represented by a nonlinear spring that incorporates
strength/stiffness degradation and pinching behavior of nonlinear sheathing-to-framing connectors.
The main structural elements of the house model correspond to non-engineered plywood/oriented strand
board shear walls with gypsum wallboard interior finish, which reflect a typical nonseismic construction
practice in Canada [16]. A generic structural representation and plan view of the UBC-SAWS model are
shown in Figure 1(a). The seismic resistance along the x-axis is weaker because of openings than that
along the y-axis. The structure is excited along the x-axis in this study. The natural vibration period of

(a) W7
(b) 0.6
Generic model & plan view
Floor 1
W6 X-axis
Normalized force with respect

0.5
W5

Floor 2
0.4
to total weight

9 ft
W4

W12 W13W14W15 W16 0.3


W3
9 ft
Floor 1 0.2
DS2 DS3 DS4
W2
0.1 DS1

20 ft
25 ft
Inverse triangle load distribution
Y W1 0.0
Nonlinear spring 0.0 0.01 0.02 0.03 0.04
X W8 W9 W10 W11
Inter-story drift ratio

Figure 1. UBC-SAWS model: (a) generic model and plan view and (b) pushover curve.

Copyright © 2014 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2015; 44:1157–1162
DOI: 10.1002/eqe
RECORD SELECTION FOR AFTERSHOCK INCREMENTAL DYNAMIC ANALYSIS 1159

the house model along the x-axis was measured as about 0.35 s in experiments [16]. On the basis of this,
spectral acceleration at 0.3 s is adopted as IM in IDA. Figure 1(b) shows a nonlinear static pushover
curve along the x-axis. The result shown in Figure 1(b) corresponds to the first-story level of the house,
which is subjected to greater seismic deformation (in fact, the house typically fails in a soft story
mechanism). Later in this study, aftershock IDA is carried out for four damage states due to a
mainshock, DS1 to DS4; inter-story drift ratios corresponding to these damage severities are indicated in
Figure 1(b). From the viewpoint of post-earthquake building inspection, DS1, DS2-DS3, and DS4 may
be associated with green-tag, yellow-tag, and red-tag categories, respectively.

2.2. Ground motion data


The selection of aftershock ground motions is the main issue in this article. Previous studies for
aftershock fragility often considered a set of 30 records used by Vamvatsikos and Cornell [10], and
thus this dataset can serve as a benchmark (referred to as V&C-30). However, this dataset does not
accompany aftershock records; consequently, back-to-back application of this dataset has been
considered in aftershock IDA [3, 9].
To examine characteristics of aftershock ground motions in real MS-AS sequences, a combined MS-AS
database based on the Kyoshin and Kiban-Kyoshin networks (K-NET/KiK-net) database for Japanese
earthquakes [4] and the Pacific Earthquake Engineering Research-Next Generation Attenuation
(PEER-NGA) database for worldwide shallow crustal earthquakes [5] is utilized. The database
contains 290 MS-AS sequences (two horizontal components per sequence). From this database, 90
record sequences are selected by applying the following criteria: (i) mainshock magnitude is between
6.5 and 7.7; (ii) earthquake occurs within continental crust (i.e., interface and in-slab events are
excluded); (iii) distance is less than 40 km; and (iv) average shear-wave velocity in the uppermost 30 m
is between 150 and 400 m/s. These criteria are adopted to resemble the record characteristics of the
V&C-30 dataset. The mainshock records of the selected 90 sequences are referred to as MS-90 (note:
mainshocks for individual sequences are identified as events with the largest magnitude), while two
sets of major aftershock records are developed for use in aftershock IDA. The AS1-90 dataset consists
of aftershock records that have the second largest magnitude within the sequences, and the AS2-90
dataset comprises aftershock records with the largest peak ground acceleration (PGA).
The magnitude-distance distributions of the four datasets are shown in Figure 2(a). Generally, the
V&C-30 and MS-90 data overlap, whereas magnitudes of the AS1-90 and AS2-90 datasets are
smaller than those of the mainshock records. The smaller magnitudes of major aftershocks are typical
features, observed in real MS-AS sequences [17] (i.e., empirical Bath’s law). Figure 2(b) compares
the response spectra statistics (median and 16th/84th percentile curves) for the four datasets. The
overall amplitudes of the MS-90 dataset are slightly greater than those of the V&C-30 dataset, but
the spectral shapes for the two datasets are very similar. On the other hand, the spectral amplitudes of

(a) (b)
8.0 5.0
V&C-30 Marker: Median
MS-90 Broken line: 16th/84th
7.5 percentile curve
AS1-90
AS2-90 1.0
Acceleration (g)

7.0
Magnitude

6.5

0.1
6.0
V&C-30
MS-90
5.5 AS1-90
AS2-90
5.0 0.01
0.1 1.0 10 100 0.01 0.1 1.0 10
Distance (km) Vibration period (s)

Figure 2. (a) Magnitude-distance distribution of four ground motion datasets and (b) 5%-damped elastic
response spectra of four ground motion datasets.

Copyright © 2014 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2015; 44:1157–1162
DOI: 10.1002/eqe
1160 K. GODA

the AS1-90 and AS2-90 datasets are less than those of the mainshocks, and their response spectral
shapes are different from those of the mainshocks. The response spectra at moderate to long periods
for the aftershocks decrease more rapidly than those for the mainshocks; this can be explained by the
different record characteristics associated with the mainshocks and aftershocks (i.e., smaller
magnitudes and slightly greater distances; Figure 2(a)). The observed features of the MS-AS data
suggest that when these records are used in IDA, nonlinear responses caused by the scaled
mainshock and aftershock records may differ significantly. This will be investigated in Section 3.

2.3. Analysis procedure and setup


Using the four datasets, aftershock IDA is conducted. Four damage states, DS1 to DS4, induced by
mainshocks are considered (Figure 1(b)). Firstly, scale factors that are required to attain intended
drift ratios are obtained through conventional IDA using mainshock records (V&C-30 or MS-90)
alone. Subsequently, following the scaled mainshocks, aftershock IDA is implemented by using the
V&C-30 (back-to-back), MS-90 (back-to-back), AS1-90, and AS2-90 datasets (note: 30 s of zeros are
inserted between mainshock and aftershock parts). The seismic intensity (i.e., spectral acceleration at
0.3 s) is varied from 0.2 to 5.0 g (40 values). In total, about four million runs of nonlinear dynamic
analyses are carried out. In IDA, the structural collapse is defined as a situation where the maximum
inter-story drift ratio of 0.08 is exceeded (note: the results presented in Section 3 are not sensitive to
the collapse limit; sensitivity of the results is examined by varying the collapse limit from 0.06 to 0.1).

3. RESULTS

The conventional mainshock IDA results are shown in Figure 3(a) using the V&C-30 and MS-90
datasets. Overall, the results are similar, which is expected from the comparison of response spectral
shapes of the record sets (Figure 2(b)). The similarity of the two cases serves as a useful reference
in interpreting the aftershock IDA results presented in the following from the viewpoint of the
benchmark record dataset selected by [10].
Figure 3(b) compares the mainshock IDA curves with the aftershock IDA on the basis of back-to-
back analysis using the real mainshock dataset (i.e., MS-90). For the aftershock IDA, the mainshock
damage state DS4 is considered (i.e., a mainshock induces the maximum inter-story drift ratio of
0.03; Figure 1(b)). The aftershock IDA curves include the effects of mainshocks; thus, the curves
sharply increase at low seismic intensity levels (i.e., no additional damage). At high seismic intensity
levels, the inter-story drift ratios start to increase rapidly (i.e., aftershock damage occurrence). It is
important to notice that the aftershock IDA curves are below the mainshock IDA curves, indicating
that the back-to-back application of scaled mainshocks leads to reduction of (residual) seismic

(a) (b)
4.0 4.0
Spectral acceleration at 0.3 s (g)

Spectral acceleration at 0.3 s (g)

Mainshock IDA Mainshock IDA


3.5 V&C-30: median 3.5 Aftershock IDA (BtoB)
16th/84th percentile curves
3.0 MS-90: median 3.0
16th/84th percentile curves
2.5 2.5

2.0 2.0

1.5 1.5

1.0 1.0

0.5 0.5
DS4
0.0 0.0
0.0 0.02 0.04 0.06 0.08 0.0 0.02 0.04 0.06 0.08
Maximum inter-story drift ratio Maximum inter-story drift ratio

Figure 3. (a) Mainshock IDA curves based on the V&C-30 and MS-90 datasets and (b) comparison of
mainshock IDA curves and aftershock IDA (back-to-back) curves.

Copyright © 2014 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2015; 44:1157–1162
DOI: 10.1002/eqe
RECORD SELECTION FOR AFTERSHOCK INCREMENTAL DYNAMIC ANALYSIS 1161

capacity. This result is in agreement with previous studies [1, 3, 9]. When DS1 and DS2 are considered
(results are not shown for brevity), the mainshock IDA and aftershock IDA results are almost identical,
indicating that minor-to-moderate damage due to mainshocks does not affect the residual seismic
capacity against collapse for this structural model (note: the results should not be generalized to other
structures). For this reason, in the following, results for DS4 are mainly discussed.
Figure 4(a) compares the aftershock IDA results using the real MS-AS sequences, i.e., MS-90
(back-to-back), AS1-90, and AS2-90. It is noteworthy that at high seismic excitation levels,
nonlinearity of the structural model is severe because of stiffness degradation and strength
deterioration, and consequently, vibration period of the damaged structure becomes longer than that
of the intact structure. This means that spectral content of input ground motions at longer vibration
periods than the fundamental vibration period has more influence on the structural response.
Figure 4(a) clearly shows that the aftershock IDA curves based on AS1-90 and AS2-90 (which are
similar for this case) are significantly different from those based on MS-90 (back-to-back). The main
reason for requiring significantly greater spectral acceleration is that generally real aftershock
records (both magnitude-based and PGA-based definitions) have much less long-period spectral
content, in comparison with mainshock records (Figure 2(b)). For instance, when the mainshock and
aftershock records are scaled to the same spectral acceleration level at 0.3 s, spectral accelerations at
1.0 to 2.0 s differ by a factor of 2 to 3. Moreover, the results shown in Figure 4(a) should not be
misinterpreted as increased residual seismic capacity; definitely, major aftershocks cause further
damage to the mainshock-damaged structure. Simply, the aftershock IDA curves based on back-to-back
application of mainshocks and based on realistic aftershock records cannot be compared in such a
format. In other words, residual seismic capacities cannot be measured on the basis of the adopted
seismic intensity parameter (i.e., Sa(T1)). Finally, Figure 4(b) shows the comparison of the probability
of collapse as a function of seismic intensity measure. This figure illustrates that the record
characteristics of aftershock ground motions (which are scaled in IDA) can have a major impact on
the collapse vulnerability assessment.

4. DISCUSSION

The response spectral shape of ground motions is an important feature for record selection, when the
scaled records are used in nonlinear dynamic analysis. On the basis of an extensive real MS-AS
database, mainshock magnitudes are on average larger than magnitudes of major aftershocks
within the same sequences. This leads to different response spectral shapes for mainshocks and
major aftershocks. A numerical example presented in this article demonstrated that the effects of
aftershock record selection can have significant influence on the aftershock fragility and collapse
vulnerability assessment. Qualitatively, the observations regarding the impact of aftershock record

(a) (b)
4.0 1.0
Spectral acceleration at 0.3 s (g)

MS-90-BtoB DS4
3.5 0.9
AS1-90
0.8
Probability of collapse

3.0 AS2-90
0.7
2.5 0.6
2.0 0.5

1.5 0.4
0.3
1.0
0.2 MS-90-BtoB
0.5 AS1-90
0.1
DS4 AS2-90
0.0 0.0
0.0 0.02 0.04 0.06 0.08 0.0 1.0 2.0 3.0 4.0 5.0
Maximum inter-story drift ratio Spectral acceleration at 0.3 s (g)

Figure 4. Comparison of aftershock IDA curves (a) and collapse probability curves (b) by considering the
MS-90 (back-to-back), AS1-90, and AS2-90 datasets (for DS4).

Copyright © 2014 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2015; 44:1157–1162
DOI: 10.1002/eqe
1162 K. GODA

selection on the nonlinear response estimation are applicable to other cases. However, quantitatively,
the results should not be generalized. This is because the response spectral shape effects depend on
various factors, such as dynamic characteristics (intact vibration period and higher-mode periods),
hysteretic behavior, structural redundancy, and collapse mechanism/mode. It is noted that the
assessment that was presented adopts spectral acceleration at the fundamental vibration period as IM
(on which record scaling is based), while use of advanced seismic intensity measures is likely to
suppress the response spectral shape effects [13, 14]. In either case, it is desirable to use ground
motion input that is representative for seismic environments under consideration. For instance, in
major subduction zones (e.g., Japan and Cascadia), earthquakes having different characteristics
contribute to overall seismic hazard, and thus the selection should reflect record features that are
attributed to physical environments (e.g., frequency content and duration) [8]. This will reduce the
possibility of having biased assessment of the inelastic seismic demand parameters (note: different
seismic demand parameters are sensitive to different record features).

ACKNOWLEDGEMENTS
The work is supported by the Alexander von Humboldt Foundation. Ground motion data for Japanese earth-
quakes and worldwide crustal earthquakes were obtained from the K-NET/KiK-net databases at http://www.
kyoshin.bosai.go.jp/ and the PEER-NGA database at http://peer.berkeley.edu/nga/index.html, respectively.
The author is grateful to constructive comments and suggestion provided by three anonymous reviewers.

REFERENCES
1. Luco N, Bazzurro P, Cornell CA. Dynamic versus static computation of the residual capacity of a mainshock-
damaged building to withstand an aftershock. 13th World Conference on Earthquake Engineering, Vancouver, Canada,
Paper 2405, 2004.
2. Li Q, Ellingwood BR. Performance evaluation and damage assessment of steel frame buildings under mainshock–
aftershock earthquake sequence. Earthquake Engineering and Structural Dynamics 2007; 36(3):405–427.
3. Ryu H, Luco N, Uma SR, Liel AB. Developing fragilities for mainshock-damaged structures through incremental
dynamic analysis. 9th Pacific Conference on Earthquake Engineering, Auckland, New Zealand, Paper 225, 2011.
4. Goda K. Nonlinear response potential of mainshock–aftershock sequences from Japanese earthquakes. Bulletin of the
Seismological Society of America 2012; 102(5):2139–2156.
5. Goda K, Taylor CA. Effects of aftershocks on peak ductility demand due to strong ground motion records from
shallow crustal earthquakes. Earthquake Engineering & Structural Dynamics 2012; 41(15):2311–2330.
6. Ruiz-García J. Mainshock-aftershock ground motion features and their influence in building’s seismic response.
Journal of Earthquake Engineering 2012; 16(5):719–737.
7. Ebrahimian H, Jalayer F, Asprone D, Lombardi AM, Marzocchi W, Prota A, Manfredi G. A performance-based
framework for adaptive seismic aftershock vulnerability assessment. Earthquake Engineering & Structural Dynam-
ics 2014. DOI: 10.1002/eqe.2444.
8. Goda K, Salami R. Inelastic seismic demand estimation of wood-frame houses subjected to mainshock-aftershock
sequences. Bulletin of Earthquake Engineering 2014; 12(2):855–874.
9. Raghunandan M, Liel AB, Luco N. Aftershock collapse vulnerability assessment of reinforced concrete frame
structures. Earthquake Engineering & Structural Dynamics 2014; in review.
10. Vamvatsikos D, Cornell CA. Direct estimation of the seismic demand and capacity of oscillators with multi-linear
static pushovers through IDA. Earthquake Engineering & Structural Dynamics 2006; 35(9):1097–1117.
11. Luco N, Bazzurro P. Does amplitude scaling of ground motion records result in biased nonlinear structural drift
responses? Earthquake Engineering & Structural Dynamics 2007; 36(13):1813–1835.
12. PEER Ground Motion Selection and Modification (GMSM) Working Group. Evaluation of ground motion selection
and modification methods: predicting median interstory drift response of buildings. PEER Report 2009/01, University
of California at Berkeley, Berkeley, CA, 2009.
13. Luco N, Cornell CA. Structure-specific scalar intensity measures for near-source and ordinary earthquake ground
motions. Earthquake Spectra 2007; 23(2):357–392.
14. Tothong P, Luco N. Probabilistic seismic demand analysis using advanced ground motion intensity measures.
Earthquake Engineering & Structural Dynamics 2007; 36(13):1837–1860.
15. Folz B, Filiatrault A. Seismic analysis of woodframe structures I: model formulation. Journal of Structural Engineering
2004; 130(9):1353–1360.
16. White TW, Ventura CE. Seismic performance of wood-frame residential construction in British Columbia. Technical
Report, Earthquake Eng. Research Facility Report No. 06-03, University of British Columbia, Vancouver, Canada,
2006.
17. Shcherbakov R, Turcotte DL, Rundle JB. Aftershock statistics. Pure and Applied Geophysics 2005; 162(6-7):1051–1076.

Copyright © 2014 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2015; 44:1157–1162
DOI: 10.1002/eqe

You might also like