You are on page 1of 385

SEISMIC COLLAPSE RISK ASSESSMENT OF BUILDINGS: EFFECTS OF

INTENSITY MEASURE SELECTION AND COMPUTATIONAL APPROACH

A DISSERTATION
SUBMITTED TO THE DEPARTMENT OF CIVIL AND
ENVIRONMENTAL ENGINEERING
AND THE COMMITTEE ON GRADUATE STUDIES
OF STANFORD UNIVERSITY
IN PARTIAL FULFILLMENT OF THE REQUIREMENTS
FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY

Laura Eads
October 2013
2013 by Laura Catherine Eads. All Rights Reserved.
Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons Attribution-


Noncommercial 3.0 United States License.
http://creativecommons.org/licenses/by-nc/3.0/us/

This dissertation is online at: http://purl.stanford.edu/wp898br0348

ii
I certify that I have read this dissertation and that, in my opinion, it is fully adequate
in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Eduardo Miranda, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequate
in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Jack Baker

I certify that I have read this dissertation and that, in my opinion, it is fully adequate
in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Dimitrios Lignos

Approved for the Stanford University Committee on Graduate Studies.


Patricia J. Gumport, Vice Provost for Graduate Education

This signature page was generated electronically upon submission of this dissertation in
electronic format. An original signed hard copy of the signature page is on file in
University Archives.

iii
Abstract

Life safety and collapse prevention have always been primary goals of earthquake
engineering, but the collapse risk of structures has not been explicitly quantified until
recently. Advances in computational power and the development of models that can more
accurately reproduce the behavior of structural components through failure have made
collapse risk assessment possible. Interest in quantifying the collapse risk has risen
significantly with the advent of performance-based earthquake engineering (PBEE),
which considers uncertainties in the seismic hazard and structural response and seeks to
engineer structures so that they achieve a desired level of performance in terms of
expected monetary losses, downtime and casualties. In fact, collapse risk assessment is a
necessary part of the PBEE process as collapse contributes to the expected economic
losses, downtime and casualties resulting from a seismic event.

Collapse risk assessment of a structure involves combining information about the seismic
hazard at the site with the behavior of the structure. Typically an intensity measure (IM)
is used to describe the level of ground motion shaking and quantify the seismic hazard.
The behavior of the structure is then characterized through nonlinear response history
analysis (RHA) by subjecting the structure to many different ground motions at various
intensity levels. A collapse fragility curve, which describes how the probability of
collapse increases as a function of the ground motion intensity, is constructed based on
analysis results and combined with the seismic hazard curve to compute the mean annual
frequency of collapse (c), a powerful collapse risk metric that describes the structures
expected frequency of collapse in a year.

This dissertation focuses on evaluating the effects of IM selection and computational


approach on the computed collapse risk, and contributions are made in the following
areas:
Uncertainty in the collapse fragility curve and c due to the number of ground
motions used in RHA is explicitly quantified and evaluated.

iv
A computationally efficient method for predicting c is proposed that limits RHAs
to two intensity levels, and the performance of this method is illustrated using a
case study.
The ability of different IMs to efficiently predict structural response is evaluated,
and reasons why certain IMs are more efficient than others are provided.
The performance of different IMs is evaluated with respect to the reliability of the
collapse risk estimate. Reasons why certain IMs give more reliable estimates are
provided.
Having been identified as a good IM for collapse risk assessment, the spectral
acceleration averaged over a period range (Saavg) is studied extensively using
various structures. Recommendations for period ranges that lead to the maximum
efficiency of this IM in predicting collapse are made based on structural
properties. The effects of the number of periods used and the spacing of these
periods within the period range on the efficiency are also considered.

v
Acknowledgements

I would like to thank my advisor, Professor Eduardo Miranda, for his guidance, advice,
and support throughout my graduate studies. I would also like to thank him for his
encouragement, especially over the past year while completing my dissertation. I have
learned a lot from working with him and am grateful for all the attention and time he
devoted to me and my research. I also really appreciated his colorful analogies and the
PS lines of his emails, which often contained bits of encouragement, humor, and/or
comments about the Notre Dame football team.

I would also like to thank Professor Dimitrios Lignos, who served as a co-advisor for this
work. His guidance, advice, support, and encouragement are very much appreciated. I am
especially grateful for his help involving structural modeling and computer programming,
including his assistance with learning OpenSees, and for sharing his solution algorithm
and incremental dynamic analysis computer codes. I really enjoyed working with him,
especially in the early years when we were both in the Blume Center.

The late Professor Helmut Krawinkler also acted as a co-advisor, and I feel very fortunate
that I had the opportunity to work with him and to learn from his wisdom. Though he was
not able to see the completion of this work, his guidance and influence were strongly felt.

Professor Jack Baker also deserves a big thank you for serving on my qualification, oral
defense, and reading committees. I really appreciate his advice and thoughtful comments
that really helped improve the quality of this work. I would also like to thank Professor
Greg Deierlein for being a part of my oral defense committee and Professor Aaron Gitler
for serving as the committee chair.

Thanks to Professor Farzin Zareian for sharing the results from his extensive analysis on
the collapse of generic moment-resisting frame and shear wall structures. Thanks also go
to Professor Curt Haselton for providing and maintaining a website with the results of his

vi
analysis on the collapse of modern reinforced concrete moment-resisting frame
structures. Data from these sources was used throughout the second half of this
dissertation.

I would also like to thank Eve Martinez, Racquel Hagen, Kim Vonner, Jill Nomura, and
Brenda Sampson for their administrative support and assistance throughout my graduate
studies.

This work was funded by a number of sources, including the Shah Family Fellowship, the
John A. Blume Fellowship, and the Network for Earthquake Engineering Simulation
(NEES) program of the National Science Foundation. This support is gratefully
appreciated and acknowledged.

The students in the Blume Center have helped make my journey through graduate school
a great experience. I learned a lot through our academic interactions and discussions, and
I also enjoyed plenty of our non-academic interactions. In particular, I am grateful for the
special bonds Ive formed with both former and current members of the Miranda research
group. I am also fortunate to have had great officemates throughout the years that have
made day-to-day life in Office 214 very enjoyable.

Finally, I would like to thank my parents and my sister for their support and
encouragement throughout these years. They were truly invaluable.

vii
Contents

Abstract iv

Acknowledgements vi

1 Introduction 1
1.1 Motivation and background ................................................................................... 1
1.2 Objectives and scope ............................................................................................. 2
1.3 Organization .......................................................................................................... 4

2 Assessing Collapse Risk 6


2.1 Overview.............................................................................................................. 6
2.2 Previous analytical studies on structural collapse ............................................... 6
2.3 Methods for evaluating collapse risk ................................................................... 19
2.4 Evaluation of different collapse risk metrics ....................................................... 23
2.4.1 Probability of collapse at a specific ground motion intensity ................... 23
2.4.2 Median collapse intensity ......................................................................... 27
2.4.3 Collapse margin ratio ................................................................................ 28
2.4.4 Mean annual frequency of collapse (c) .................................................... 29
2.4.5 Probability of collapse over a given time period ...................................... 33
2.5 Simplified collapse risk assessment..................................................................... 34
2.6 Summary .............................................................................................................. 37

3 Collapse Risk Assessment Case Study: 4-Story Steel Frame Structure 38


3.1 Overview............................................................................................................. 38
3.2 Site and seismic hazard analysis ......................................................................... 38
3.3 Ground motion selection..................................................................................... 41
3.4 Description of structure and modeling ............................................................... 43
3.4.1 General description of structure ............................................................... 43

viii
3.4.2 General description of model ................................................................... 44
3.4.3 Mass and loading ..................................................................................... 45
3.4.4 Damping................................................................................................... 45
3.4.5 Hysteretic behavior .................................................................................. 46
3.4.6 Definition of collapse ............................................................................... 48
3.5 Simulation and collapse risk results ................................................................... 50
3.6 Influence of the seismic hazard curve ................................................................ 55

4 Statistical Uncertainty in the Collapse Fragility and Mean Annual Frequency


of Collapse due to the Number of Ground Motions Considered 63
4.1 Overview............................................................................................................. 63
4.2 Background and literature review ....................................................................... 64
4.3 Confidence intervals on the probability of collapse conditioned on intensity.... 69
4.4 Case study of statistical uncertainty in the collapse risk .................................... 73
4.4.1 Confidence intervals on the collapse fragility curve................................ 74
4.4.2 Confidence intervals on the mean annual frequency of collapse ............. 75
4.5 Proposed method for estimating the mean annual frequency of collapse .......... 76
4.5.1 Method overview ..................................................................................... 77
4.5.2 Description of method.............................................................................. 77
4.5.3 Discussion of method ............................................................................... 79
4.5.4 Evaluation of the method via a case study ............................................... 80
4.6 Summary ............................................................................................................. 95

5 Spectral Shape and Structural Response 97


5.1 Overview ............................................................................................................ 97
5.2 Literature review ................................................................................................ 97
5.3 What causes large displacement increments? ....................................................104
5.4 Factors influencing spectral shape .....................................................................108
5.4.1 Response to individual pulses ..................................................................108
5.4.1.1 Using at-rest initial conditions ..................................................110
5.4.1.2 Effect of initial conditions .....................................................114

ix
5.4.2 Response to a ground motion...................................................................118
5.4.2.1 Peak elastic response of SDOF systems with similar periods ....119
5.4.2.2 Formation of spectral peaks and valleys ....................................122
5.4.2.3 Elastic versus inelastic response.................................................124
5.5 Spectral shape of damaging records ...................................................................126
5.6 Measures of spectral shape .................................................................................138
5.7 Implications for collapse risk assessment ..........................................................150
5.7.1 Influence of SaRatio on collapse risk assessment....................................152
5.7.2 Considering SaRatio in collapse risk assessment ....................................153
5.7.2.1 Full IDA method ........................................................................157
5.7.2.2 Limited stripe analysis................................................................159
5.8 Summary .............................................................................................................163

6 Intensity Measures for Predicting Collapse 168


6.1 Overview.............................................................................................................168
6.2 Background and literature review .......................................................................169
6.3 Spectral acceleration averaged over a period range (Saavg) as an IM .................180
6.3.1 Relationship between Saavg,col, Sa(T1)col and SaRatio...............................181
6.3.2 Why Saavg is more efficient than Sa(T1) in predicting the collapse
intensity ....................................................................................................184
6.3.3 Evaluating the sufficiency of Saavg ..........................................................188
6.3.3.1 Magnitude .........................................................................................190
6.3.3.2 Distance ............................................................................................195
6.3.3.3 Spectral shape measures ...................................................................196
6.3.3.3.1 ...............................................................................................197
6.3.3.3.2 SaRatio ....................................................................................198
6.3.4 Obtaining a hazard curve for Saavg...................................................................................... 200
6.4 Evaluation of different IMs for collapse risk assessment ...................................204
6.4.1 Effects of efficiency and sufficiency .......................................................205
6.4.2 Effect of ground motion selection............................................................215
6.4.2.1 Description of ground motion sets .............................................216

x
6.4.2.2 Collapse risk comparisons ..........................................................219
6.5 Summary .............................................................................................................223

7 Average Spectral Acceleration (Saavg): Optimal Period Ranges for


Various Structures for Maximum Efficiency 227
7.1 Overview.............................................................................................................227
7.2 Literature review .................................................................................................227
7.3 Effect of period spacing and number of periods on the value of Saavg ...............229
7.4 Effect of period range used to compute Saavg on the efficiency .........................231
7.4.1 Case study structure .................................................................................232
7.4.1.1 Arithmetic period spacing ..........................................................232
7.4.1.2 Logarithmic period spacing ........................................................241
7.4.2 Generic SDOF systems ............................................................................244
7.4.3 Generic moment-resisting frame structures .............................................248
7.4.3.1 Arithmetic period spacing ..........................................................248
7.4.3.2 Logarithmic period spacing ........................................................255
7.4.3.3 Comparison of period spacing schemes .....................................259
7.4.4 Modern reinforced concrete moment-resisting frame structures .............261
7.4.4.1 Arithmetic period spacing ..........................................................262
7.4.4.2 Logarithmic period spacing ........................................................268
7.4.4.3 Comparison of period spacing schemes .....................................271
7.4.5 Generic shear wall structures ...................................................................273
7.4.5.1 Arithmetic period spacing ..........................................................274
7.4.5.2 Logarithmic period spacing ........................................................281
7.4.5.3 Comparison of period spacing schemes .....................................284
7.5 Summary and discussion ....................................................................................286

8 Summary and Conclusions 296


8.1 Overview.............................................................................................................296
8.2 Summary of findings and conclusions ................................................................297
8.2.1 Collapse risk assessment ..........................................................................297

xi
8.2.2 Uncertainty in the collapse fragility and mean annual frequency of
collapse ....................................................................................................298
8.2.3 Spectral shape ..........................................................................................300
8.2.4 Intensity measures for predicting collapse ...............................................302
8.2.5 Effect of period range on the efficiency of Saavg .....................................304
8.3 Limitations and suggestions for future work ......................................................307
8.4 Implications for engineering practice ................................................................310
8.5 Concluding remarks ............................................................................................310

A MRCD 137 Ground Motion Set 312

B Description of Case Studies Used in Chapters 5-7 322


B.1 Generic moment-resisting frame structures .......................................................322
B.2 Generic shear wall structures .............................................................................323
B.3 Reinforced concrete moment-resisting frame structures ....................................323

C CS Ground Motion Sets Used in Chapters 6 and 7 325


C.1 CS, (1) = 1.7 ground motion set ......................................................................326
C.2 CS, (1) = 1.8 ground motion set ......................................................................329

xii
List of Tables

Table 3.1 Modeling parameters for nonlinear rotational springs used in case study ..... 48
Table 3.2 Locations used in case study .......................................................................... 56
Table 3.3 c values for case study .................................................................................. 58
Table 4.1 Computational effort and level of accuracy when estimating c using the
proposed and IDA methods ........................................................................... 89
Table 5.1 Summary of correlation between collapse intensity and or SaRatio ...........146
Table 5.2 Collapse risk assessment results using different record sets ..........................152
Table 6.1 Summary of dispersion on the collapse fragility curve for Saavg and Sa(T1) .187
Table 6.2 Summary of p-values for the relationship between Saavg,col and earthquake
magnitude .......................................................................................................193
Table 6.3 Summary of p-values for the relationship Saavg,col and distance ....................196
Table 6.4 Summary of p-values for the relationship between Saavg,col and .................198
Table 6.5 Summary of p-values for the relationship between Saavg,col and SaRatio ......199
Table 6.6 Collapse risk metrics for different IMs with the MRCD 137 ground
motion set .......................................................................................................207
Table A.1 Summary of MRCD 137 set ..........................................................................314
Table A.2 Records comprising the MRCD 137 ground motion set ................................314
Table C.1 Records comprising the CS, (T1) = 1.7 ground motion set ..........................327
Table C.2 Records comprising the CS, (T1) = 1.8 ground motion set ..........................331

xiii
List of Figures

Figure 2.1 Hypothetical collapse fragility curves for two structures located at the
same site ...................................................................................................... 25
Figure 2.2 Hypothetical structure located at two different sites: (a) collapse fragility
curve; (b) seismic hazard curves that cross at the 2/50 hazard level .......... 26
Figure 2.3 Illustration of c deaggregation process: (a) collapse fragility curve;
(b) numerical derivative of seismic hazard curve; (c) c deaggregation ..... 31
Figure 2.4 Illustration of identifying the contribution to c from a range of
intensities: (a) collapse fragility curve; (b) c deaggregation ..................... 32
Figure 3.1 Map of Bulk Mail site and surrounding faults. Faults controlling the site
hazard at T = 1 s are shown in blue (strike-slip) and red (reverse or
reverse/oblique) while other faults contributing to the hazard are shown
in white. Thin yellow lines denote freeways. ............................................. 39
Figure 3.2 Probabilistic seismic hazard deaggregation for Bulk Mail Site for
Sa(T = 1 s) with a 2% probability of exceedance in 50 years .................... 40
Figure 3.3 Seismic hazard curve for the Bulk Mail Center site ................................... 41
Figure 3.4 Case-study structure: (a) plan view; (b) elevation view along gridline 4 ... 44
Figure 3.5 Modified Ibarra-Medina-Krawinkler deterioration model and definition
of parameters ............................................................................................... 47
Figure 3.6 First-story drift ratio (IDR1) history under a Chi-Chi ground motion for
various ground motion intensities, illustrating how the IDR limit affects
the collapse intensity ................................................................................... 49
Figure 3.7 Pushover analysis of case study structure: (a) schematic of collapse
mechanism; (b) pushover curve .................................................................. 50
Figure 3.8 IDA curves for each of the 274 records used in the case study. The thick
black line denotes the median Sa(T1) conditioned on the value of
maximum IDR ............................................................................................ 51

xiv
Figure 3.9 Collapse fragility for the case study including individual collapse
intensities and various probability distributions fit to the data. The
lognormal distribution is selected as the best fit to the data. The
lognormal curve is defined by a median collapse intensity of
Sa(T1) = 1.03 g and a lognormal standard deviation of ln = 0.39 .............. 52
Figure 3.10 Case study collapse risk assessment: (a) collapse fragility curve;
(b) slope of the seismic hazard curve; (c) c deaggregation curve .............. 55
Figure 3.11 Seismic hazard curves for Sa(T1) for various locations in the US .............. 57
Figure 3.12 Case study collapse risk assessment for locations throughout the US:
collapse fragility curve in (a) and (b); slopes of the seismic hazard
curves in (c) and (d); c deaggregation curves in (e) and (f). Plots on the
left (i.e., (a) (c) and (e)) pertain to the western US locations while plots
on the right (i.e., (b) (d) and (f)) pertain to the central and eastern US
locations ...................................................................................................... 59
Figure 3.13 Cumulative contribution to c as a function of intensity for: (a) western
US sites; (b) central and eastern US sites. All sites have a collapse risk
of c = 2.0x10-4 and dispersion on the collapse fragility curve of ln = 0.4 61
Figure 4.1 95% binomial confidence intervals for P(C | IM): Wilson score
confidence intervals as a function of pim for different N ........................... 72
Figure 4.2 Realizations of pim for im = 0.74 g with different N: (a) N = 10;
(b) N = 40; (c) N = 274 ............................................................................... 74
Figure 4.3 Collapse fragility curve with 95% confidence intervals on P(C | IM)
based on simulations using the case study results shown for different N ... 75
Figure 4.4 95% confidence intervals for c as a function of N ..................................... 76
Figure 4.5 Relationship between where a given intensity is located on
(a) the collapse fragility curve; (b) the c deaggregation curve;
(c) the cumulative contribution to c curve ................................................. 82

xv
Figure 4.6 Percent error in the mean c of the bootstrap samples with respect to the
true c as a function of the cumulative contribution to c used to select
the intensity levels IM1 and IM2, shown for N = 40 and an initial guess of
the collapse fragility curve of = 0.4 and: (a) = 0.5 g; (b) = 1 g;
(c) = 1.5 g ................................................................................................. 83
Figure 4.7 Error in the mean c of the bootstrap samples with respect to the true
c as a function of N for different intensity levels. The initial guesses of
the collapse fragility curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g;
(c) = 1.5 g. Note: legend entries denote the fraction of cumulative
contribution to c that were used to select IM1 and IM2, respectively. For
example, 0.8, 0.2 means that the intensity with an 80% cumulative
contribution to c was selected for IM1 and the intensity with a 20%
cumulative contribution to c was selected for IM2 .................................... 85
Figure 4.8 Width of the 95% confidence intervals on c, expressed as a percentage
of the true c, as a function of N for different intensity levels. The
initial guesses of the collapse fragility curve are = 0.4 and: (a) =
0.5 g; (b) = 1 g; (c) = 1.5 g .................................................................... 86
Figure 4.9 Comparison of error in the mean c of the bootstrap samples with respect
to the true c as a function of N for different intensity levels for the
case study collapse intensities (solid lines) and exactly lognormal
collapse intensities (dashed lines). The initial guesses of the collapse
fragility curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g; (c) = 1.5 g ...... 87
Figure 4.10 95% confidence intervals on c as a function of the number of RHAs for
the IDA and proposed methods................................................................... 88

xvi
Figure 4.11 Comparison of error in the mean c of the bootstrap samples with respect
to the true c as a function of the total number of response history
analyses used to estimate the collapse fragility curve. Results are shown
when using two or three intensity levels to estimate the collapse
fragility. Intensities with 90% and 20% cumulative contributions to c
are used for IM1 and IM2, respectively. When a third intensity is used, it
is selected as the mean of IM1 and IM2. The initial guesses of the
collapse fragility curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g;
(c) = 1.5 g ................................................................................................. 91
Figure 4.12 Width of the 95% confidence intervals on c, expressed as a percentage
of the true c, as a function of the total number of response history
analyses used to estimate the collapse fragility curve. Results are shown
when using two or three intensity levels to estimate the collapse
fragility. Intensities with 90% and 20% cumulative contributions to c
are used for IM1 and IM2, respectively. When a third intensity is used, it
is selected as the mean of IM1 and IM2. The initial guesses of the
collapse fragility curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g;
(c) = 1.5 g ................................................................................................. 92
Figure 4.13 Collapse risk comparison of case study results from Section 3.5 and
those presented by Eads et al. (2013) for: (a) seismic hazard curve; (b)
collapse fragility curve; (c) c deaggregation.............................................. 93
Figure 4.14 Comparison of error in the mean c of the bootstrap samples with respect
to the true c as a function of N for different intensity levels using the
Eads et al. (2013) case study results. The initial guesses of the collapse
fragility curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g; (c) = 1.5 g. ..... 94
Figure 5.1 TCU110E ground acceleration recording from the 1999 Chi-Chi
earthquake ...................................................................................................109
Figure 5.2 Acceleration pulse with the largest incremental velocity from the
TCU110E record and an equivalent half-sine pulse with same area and
duration .......................................................................................................110

xvii
Figure 5.3 Displacement response spectrum for the equivalent half-sine
acceleration pulse using at-rest initial conditions .......................................111
Figure 5.4 Displacement response and occurrence of peak response for selected
systems subjected to a half-sine acceleration pulse: (a) peak response
transitioning to first local minima; (b) peak response transitioning from
first local minima to end of pulse. ..............................................................112
Figure 5.5 Normalized displacement response spectrum given a half-sine
acceleration pulse and at-rest initial conditions ..........................................113
Figure 5.6 Effect of half-sine pulse amplitude on displacement demands of SDOFs
with at-rest initial conditions: (a) acceleration pulse; (b) normalized
displacement spectrum ................................................................................114
Figure 5.7 Response to a half-sine pulse with different initial conditions:
(a) displacement response of a 2.5-s system; (b) half-sine acceleration
pulse ............................................................................................................115
Figure 5.8 (a)-(e) Peak displacement as a function of pulse arrival time t for various
amplitudes of initial conditions: (a) td/T = 0.1; (b) td/T = 0.3;
(c) td/T = 0.5; (d) td/T = 0.75; (e) td/T = 1.5; (f) initial conditions of the
system as a function of pulse arrival time t.................................................117
Figure 5.9 TCU110E record: (a) displacement response spectrum; (b) Time at
Peak (T@P) spectrum ...............................................................................119
Figure 5.10 TCU110E record: (a) displacement response for systems in the spectral
displacement peak; (b) ground acceleration ...............................................121
Figure 5.11 TCU110E displacement response of systems near the spectral
displacement valley at T = 3.29 s, illustrating the peak displacement
shifting from one local maxima to another as the period changes ..............123
Figure 5.12 Elastic and inelastic displacement response spectra for the TCU110E
record ..........................................................................................................125
Figure 5.13 TCU110E record: (a) displacement history for T = 2.5 s; (b) ground
acceleration .................................................................................................126

xviii
Figure 5.14 Displacement spectra of records producing the 10 lowest and 10 highest
collapse intensities, where records are scaled to a common value of
Sd(T1) ..........................................................................................................127
Figure 5.15 TCU063N record scaled to Sd(T1) = 10 cm: (a) displacement response
spectrum; (b) T@P spectrum; (c) displacement response spectra
including spectra for the pulse with the largest incremental velocity;
(d) ground acceleration history ...................................................................128
Figure 5.16 TCU063N record: (a) 1st story drift ratio history; (b) unscaled ground
acceleration history .....................................................................................130
Figure 5.17 TCU055N record scaled to Sd(T1) = 10 cm: (a) displacement response
spectrum; (b) T@P spectrum; (c) displacement response spectra
including spectra for selected pulse; (d) ground acceleration history .....131
Figure 5.18 Pulse from the TCU055N record and an equivalent half-sine pulse:
(a) ground acceleration; (b) displacement response spectrum ....................132
Figure 5.19 Displacement spectra of the TCU055N record scaled to Sd(T1) = 10 cm,
including two pulse spectra .....................................................................133
Figure 5.20 TCU055N record: (a) 1st story drift ratio history; (b) ground acceleration
history .........................................................................................................134
Figure 5.21 Comparison of displacement spectra for records with the (a) lowest;
(b) median; and (c) highest collapse intensities, where all records are
scaled to the same Sd(T1) ............................................................................135
Figure 5.22 Comparison of spectra for records with similar spectral shapes but
significantly different collapse intensities ..................................................136
Figure 5.23 TCU057N record: (a) 1st story drift ratio history; (b) ground acceleration
history .........................................................................................................137
Figure 5.24 Comparison of pseudo-acceleration spectra for records with the
(a) lowest; (b) median; and (c) highest collapse intensities, where all
records are scaled to the same Sa(T1) .........................................................139
Figure 5.25 Pseudo-acceleration spectra for records scaled to a common value of
Sa(T1) with (a) SaRatio < 1; (b) SaRatio > 1 ..............................................140

xix
Figure 5.26 Pseudo-acceleration spectra for: (a, b) unscaled negative records;
(c, d) unscaled positive records; (e) records scaled to common value of
Sa(T1) ..........................................................................................................142
Figure 5.27 Contour plot of the correlation coefficient () between the natural
logarithm of collapse intensity and the natural logarithm of SaRatio for
the case study as a function of the period range used to compute SaRatio 143
Figure 5.28 Natural log of collapse intensity versus natural log of SaRatio ..................144
Figure 5.29 Natural log of collapse intensity versus (T1) .............................................145
Figure 5.30 Unscaled pseudo-acceleration spectra with predicted spectral shape from
Boore and Atkinson 2008 (BA08) model for selected ground motions:
(a) negative , high Sa(T1)col; (b) neutral , high Sa(T1)col; (c) neutral ,
low Sa(T1)col; (d) positive , low Sa(T1)col...................................................148
Figure 5.31 Natural log of collapse intensity versus ...................................................150
Figure 5.32 SaRatio, computed using a period range from 0.2T1 to 3T1, where
T1 = 1.33 s, as a function of magnitude and distance. SaRatio values are
computed from the CMS based on the BA08 GMPE for an unspecified
fault type and Vs30 = 285 m/s. Different values of (T1) are used to
compute the CMS: (a) (T1) = 0; (b) (T1) = 1; (c) (T1) = 2 ......................154
Figure 5.33 Logistic regression on collapse state and ln(SaRatio) for case study data
at Sa(T1) = 0.7 g ..........................................................................................161
Figure 6.1 Sa(T1)col versus SaRatio with linear regression through the origin .............181
Figure 6.2 Collapse fragility curve for (a) IM = Sa(T1); (b) IM = Saavg. The circles
are the collapse intensities of individual ground motions, and the solid
lines are the lognormal fragility curves obtained using the method of
moments ......................................................................................................183
Figure 6.3 Collapse intensity scatter plots for (a) IM = Sa(T1); (b) IM = Saavg. The
horizontal black line indicates the median value ........................................185
Figure 6.4 ln(Saavg,col) plotted against earthquake magnitude with the ordinary
linear least squares regression of the data ...................................................190

xx
Figure 6.5 Response spectra predicted by BA08 for different magnitudes with Rjb =
12 km, Vs = 285 m/s, and unspecified fault type: (a) unscaled spectra;
(b) spectra scaled to a common value of Saavg (linear Sa(T) axis);
(c) spectra scaled to a common value of Saavg (logarithmic Sa(T) axis).
The dashed vertical lines bracket the period range used to compute Saavg .192
Figure 6.6 p-values for the relationship between magnitude and ln(Saavg,col) or
ln(Sa(T1)col) for the RC MRF case study structures ....................................195
Figure 6.7 ln(Saavg,col) plotted against source-to-site distance with the ordinary
linear least squares regression of the data ...................................................196
Figure 6.8 ln(Saavg,col) plotted against (T1) with the ordinary linear least squares
regression of the data ..................................................................................197
Figure 6.9 ln(Saavg,col) plotted against ln(SaRatio) with the ordinary linear least
squares regression of the data .....................................................................199
Figure 6.10 Collapse risk components for different IMs: Seismic hazard curves for
simplified site: (a) Sa(T1) and Saavg; (b) PGV. Collapse fragility curves
for (c) Sa(T1), Sa(T1) with or SaRatio adjustment, and Saavg; (d) PGV.
c deaggregation curves for (e) Sa(T1), Sa(T1) with or SaRatio
adjustment and Saavg; (f) PGV. ...................................................................206
Figure 6.11 PSHA deaggregation by magnitude with associated (T1) values for
Sa(T1) = 0.72 g, the intensity at the peak of the c deaggregation ..............209
Figure 6.12 Collapse risk components for Sa(T1) illustrating the effects of improved
efficiency and improved GMPE: Seismic hazard curves for simplified
site in (a) and (b); Collapse fragility curves in (c) and (d); c
deaggregation curves in (e) and (f). Plots (a), (c) and (e) illustrate the
effect of improved efficiency while plots (b), (d) and (f) illustrate the
effect of improved GMPE prediction .........................................................212
Figure 6.13 Components influencing the value of lnSaAvg,BA08: (a) values of
lnSa(T),BA08; (b) contour of correlation coefficients between spectral
values at different periods ...........................................................................213

xxi
Figure 6.14 The probability of collapse conditioned on intensity level, P(C|IM),
plotted as a function of IM, the mean annual frequency of exceedance of
the IM level for Saavg and PGV ...................................................................215
Figure 6.15 Geometric mean spectrum of different ground motion sets scaled to a
common value of (a) Sa(T1); (b) Saavg; and (c) PGV. Note the vertical
line in (a) is drawn at T1 = 1.33 s and the horizontal line in (b) denotes
the common value of Saavg over the period range between 0.2T1 and
3T1..............................................................................................................219
Figure 6.16 Summary of collapse risk results for different IMs and different ground
motion sets: (a) median collapse intensity; (b) standard deviation of
collapse intensities, lnIM; (c) mean annual frequency of collapse, c, and
probability of collapse in 50 years, Pc,50 .....................................................221
Figure 7.1 Effect of period spacing scheme and number of periods used to compute
Saavg(0.2, 3) for T1 = 1.33 s: (a) arithmetic spacing; (b) logarithmic
spacing ........................................................................................................230
Figure 7.2 Contour plot of the dispersion (ln) on the collapse intensities measured
by Saavg(a, b) as a function of the a and b values used to compute Saavg
using arithmetic period spacing ..................................................................233
Figure 7.3 Dispersion of collapse intensities measured by Saavg(a, b) using
arithmetic period spacing for the special case of a = b, which is
equivalent to Sa(aT1), as a function of a. ...................................................234
Figure 7.4 Contours of collapse risk estimates measured by Saavg(a, b) as a function
of the a and b values used to compute Saavg: (a) mean annual frequency
of collapse, c, in units of 10-4; (b) probability of collapse in 50 years,
Pc,50, expressed as a percent ........................................................................236
Figure 7.5 Contours of dispersion (ln) on the collapse intensities measured by
Saavg(a, b) as a function of the a and b values used to compute Saavg
using arithmetic period spacing for different ground motion sets:
(a) LMSR-N; (b) Set One; (c) Set #1A, FN/FP; (d) Set #1A, unrotated;
(e) CS, (T1) = 1.7; (f) CS, (T1) = 1.8 ........................................................241

xxii
Figure 7.6 Contours of the dispersion (ln) of the collapse intensities as a function
of the a and b values used to compute Saavg(a, b) for various period
spacing schemes: (a) arithmetic spacing with 100 periods; (b) arithmetic
spacing with 10 periods; (c) logarithmic spacing with 100 periods;
(d) logarithmic spacing with 10 periods .....................................................243
Figure 7.7 Contours of dispersion (ln) on the collapse intensities measured by
Saavg(a, b) as a function of the a and b values used to compute Saavg with
logarithmic period spacing for different ground motion sets: (a) LMSR-
N; (b) Set One; (c) Set #1A, FN/FP; (d) Set #1A, unrotated; (e) CS, (T1)
= 1.7; (f) CS, (T1) = 1.8 .............................................................................244
Figure 7.8 Tri-linear backbone curve used to define SDOF systems. ..........................245
Figure 7.9 Contours of dispersion (ln) on the collapse intensities measured by
Saavg(a, b) as a function of the a and b values used to compute Saavg for
various SDOF systems: (a) T = 0.1 s; (b) T = 0.2 s; (c) T = 0.5 s;
(d) T = 0.75 s; (e) T = 1.0 s; (f) T = 1.5 s; (g) T = 2.0 s; (h) T = 2.5 s;
(i) T = 3.0 s; (j) T = 3.5 s; (k) T = 4.0 s .......................................................247
Figure 7.10 Contours of the dispersion (ln) of the collapse intensities as a function
of the a and b values used to compute Saavg(a, b) using arithmetic period
spacing that are representative of results for generic MRF structures:
(a) 4 story; (b) 8 story; (c) 12 story; (d) 16 story ........................................249
Figure 7.11 Histogram of a* and b* values associated with the minimum dispersion
of Saavg(a, b) using arithmetic period spacing for generic MRF
structures: (a) 4 story; (b) 8 story; (c) 12 story; (d) 16 story ......................250
Figure 7.12 Scatter plot of b* using arithmetic period spacing for all generic MRF
structures .....................................................................................................251
Figure 7.13 b* value associated with the minimum dispersion of Saavg(a, b) as a
function of T1 using arithmetic period spacing for the generic MRF
structures .....................................................................................................252

xxiii
Figure 7.14 Contour plot showing the mean value, based on the 396 generic MRF
structures, of the difference between the dispersion on the collapse
intensities computed using Saavg(a*, b*) versus Saavg(a, b) as a function
of a and b using arithmetic period spacing: (a) difference in absolute
terms; (b) difference normalized by the dispersion based on
Saavg(a*, b*). ...............................................................................................253
Figure 7.15 Scatter plot of dispersion in collapse intensities computed using Saavg
with arithmetic period spacing for the 396 generic MRF structures for
different computations of Saavg(a, b): (a) Saavg(a*, b*) vs.
Saavg(0, b:reg); (b) Saavg(a*, b*) vs. Saavg(0, 3); (c) Saavg(0, 3) vs.
Saavg(0, b:reg). Note that b:reg is the b value obtained from the
regression based on T1 shown in Figure 7.13 ..............................................254
Figure 7.16 Contours of the dispersion (ln) of the collapse intensities as a function
of the a and b values used to compute Saavg(a, b) for logarithmic period
spacing that are representative of results for generic MRF structures:
(a) 4 story; (b) 8 story; (c) 12 story; (d) 16 story ........................................256
Figure 7.17 Histogram of a* and b* values associated with the minimum dispersion
of Saavg(a, b) using logarithmic period spacing for generic MRF
structures: (a) 4 story; (b) 8 story; (c) 12 story; (d) 16 story ......................257
Figure 7.18 Relationship between the a* and b* values associated with the minimum
dispersion of Saavg(a, b) obtained using logarithmic period spacing and
T1 for the generic MRF structures: (a) a* and T1; (b) b* and T1 .................258
Figure 7.19 Scatter plot of dispersion in collapse intensities computed using Saavg
with logarithmic period spacing for the 396 generic MRF structures for
different computations of Saavg(a, b): (a) Saavg(a*, b*) vs.
Saavg(a:reg, b:reg); (b) Saavg(a*, b*) vs. Saavg(0.2, 9); (c) Saavg(0.2, 9) vs.
Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a and b values,
respectively, obtained from the regression based on T1 shown in
Figure 7.18 ..................................................................................................259

xxiv
Figure 7.20 Comparison of the overall minimum dispersion, lnSaAvg(a*,b*), achieved
with logarithmic versus arithmetic period spacing for the 396 generic
MRF structures............................................................................................260
Figure 7.21 Comparison of the dispersion, lnSaAvg(a,b), achieved with logarithmic
versus arithmetic period spacing for the 396 generic MRF structures
using: (a) Saavg(0.2, 9) for logarithmic spacing versus Saavg(0, 3) for
arithmetic spacing; (b) Saavg(0.2, 9) for logarithmic spacing versus
Saavg(0, b:reg) for arithmetic spacing. Note that b:reg is the b value
obtained from the regression based on T1 shown in Figure 7.13 ................261
Figure 7.22 Contours of the dispersion (ln) of the collapse intensities as a function
of the a and b values used to compute Saavg(a, b) using arithmetic period
spacing that are representative of results for RC MRF structures: (a) 1
story; (b) 2 story; (c) 4 story; (d) 8 story; (e) 12 story; (f) 20 story ............263
Figure 7.23 Scatter plot of a* and b* using arithmetic period spacing for all RC
MRF structures............................................................................................264
Figure 7.24 b* value associated with the minimum dispersion of Saavg(a, b) as a
function of T1 using arithmetic period spacing for the RC MRF
structures .....................................................................................................265
Figure 7.25 Contour plot showing the mean value, based on the 30 RC MRF
structures, of the difference between the dispersion on the collapse
intensities computed using Saavg(a*, b*) versus Saavg(a, b) as a function
of a and b using arithmetic period spacing: (a) difference in absolute
terms; (b) difference normalized by the dispersion based on Saavg(a*, b*) 266
Figure 7.26 Scatter plot of dispersion in collapse intensities computed using Saavg for
the 30 RC MRF structures for different computations of Saavg(a, b) using
arithmetic period spacing: (a) Saavg(a*, b*) vs. Saavg(0, b:reg);
(b) Saavg(a*, b*) vs. Saavg(0, 3.5); (c) Saavg(0, 3.5) vs. Saavg(0, b:reg).
Note that b:reg is the b value obtained from the regression based on T1
shown in Figure 7.24...................................................................................268

xxv
Figure 7.27 Contours of the dispersion (ln) of the collapse intensities as a function
of the a and b values used to compute Saavg(a, b) using logarithmic
period spacing that are representative of results for RC MRF structures:
(a) 1 story; (b) 2 story; (c) 4 story; (d) 8 story; (e) 12 story; (f) 20 story ...269
Figure 7.28 Relationship between the a* and b* values associated with the minimum
dispersion of Saavg(a, b) obtained using logarithmic period spacing and
T1 for the 30 RC MRF structures: (a) a* and T1; (b) b* and T1 ..................270
Figure 7.29 Comparison of the dispersion, lnSaAvg(a,b), achieved with logarithmic
period spacing for the 30 RC MRF structures for different computations
of Saavg(a, b) using arithmetic period spacing: (a) Saavg(a*, b*) vs.
Saavg(a:reg, b:reg); (b) Saavg(a*, b*) vs. Saavg(0.4, 9); (c) Saavg(0.4, 9) vs.
Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a and b values
obtained from the regressions based on T1 shown in Figure 7.28 ...............271
Figure 7.30 Comparison of the dispersion, lnSaAvg(a,b), achieved with logarithmic
versus arithmetic period spacing for the 30 RC MRF structures using: (a)
the overall minimum dispersion; (b) period-independent (rigid)
recommendations of Saavg(0.4, 9) for logarithmic spacing versus
Saavg(0, 3.5) for arithmetic spacing; (c) regression-based values of
Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a and b values,
respectively, obtained from the regressions based on T1 shown in Figure
7.24 for arithmetic spacing and in Figure 7.28 for logarithmic spacing.
Also note that for arithmetic spacing a:reg = 0 ..........................................273
Figure 7.31 Contours of the dispersion (ln) of the collapse intensities as a function
of the a and b values used to compute Saavg(a, b) using arithmetic period
spacing that are representative of results for generic shear wall
structures: (a) 4 story; (b) 8 story; (c) 12 story; (d) 16 story ......................275
Figure 7.32 Histogram of a* and b* values associated with the minimum dispersion
of Saavg(a, b) for generic shear wall structures using arithmetic period
spacing: (a) 4 story; (b) 8 story; (c) 12 story; (d) 16 story..........................276
Figure 7.33 Scatter plot of b* for all generic shear wall structures using arithmetic
period spacing .............................................................................................277

xxvi
Figure 7.34 b* value associated with the minimum dispersion of Saavg(a, b) using
arithmetic period spacing as a function of T1 for the generic shear wall
structures .....................................................................................................278
Figure 7.35 Contour plot showing the mean value, based on the 252 generic shear
wall structures, of the difference between the dispersion on the collapse
intensities computed using Saavg(a*, b*) versus Saavg(a, b) as a function
of a and b for arithmetic period spacing: (a) difference in absolute terms;
(b) difference normalized by the dispersion based on Saavg(a*, b*) ...........279
Figure 7.36 Scatter plot of dispersion in collapse intensities computed using Saavg
with arithmetic period spacing for the 252 generic shear wall structures
for different computations of Saavg(a, b): (a) Saavg(a*, b*) vs.
Saavg(0, b:reg); (b) Saavg(a*, b*) vs. Saavg(0, 6); (c) Saavg(0, 6) vs.
Saavg(0, b:reg). Note that b:reg is the b value obtained from the
regression based on T1 shown in Figure 7.34 ..............................................281
Figure 7.37 Contours of the dispersion (ln) of the collapse intensities as a function
of the a and b values used to compute Saavg(a, b) with logarithmic period
spacing that are representative of results for generic shear wall
structures: (a) 4 story; (b) 8 story; (c) 12 story; (d) 16 story ......................282
Figure 7.38 Relationship between the a* and b* values associated with the minimum
dispersion of Saavg(a, b) obtained using logarithmic period spacing and
T1 for the 252 generic shear wall structures: (a) a* and T1; (b) b* and T1 ..283
Figure 7.39 Scatter plot of the dispersion in collapse intensities computed using Saavg
with logarithmic period spacing for the 252 generic shear wall structures
for different computations of Saavg(a, b): (a) Saavg(a*, b*) vs.
Saavg(a:reg, b:reg); (b) Saavg(a*, b*) vs. Saavg(0.7,15); (c) Saavg(0.7,15)
vs. Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a and b values,
respectively, obtained from the regression based on T1 shown in
Figure 7.38 ..................................................................................................284

xxvii
Figure 7.40 Comparison of the dispersion, lnSaAvg(a,b), achieved with logarithmic
versus arithmetic period spacing for the 252 generic shear wall
structures using: (a) the overall minimum dispersion; (b) period-
independent (rigid) recommendations of Saavg(0.7, 15) for logarithmic
spacing versus Saavg(0, 6) for arithmetic spacing; (c) regression-based
values of Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a and b
values, respectively, obtained from the regressions based on T1 shown in
Figure 7.34 for arithmetic spacing and in Figure 7.38 for logarithmic
spacing. Also note that for arithmetic spacing a:reg = 0 ............................285
Figure A.1 Magnitudes and distances of records in the MRCD 137 set .......................313
Figure A.2 Response spectra of records in the MRCD 137 set with the median
spectrum denoted by the thick black line ....................................................313
Figure C.1 Comparison of the target CS and the selected ground motions:
(a) geometric mean spectral values; (b) logarithmic standard deviation of
spectral values .............................................................................................326
Figure C.2 Response spectra of ground motions in the CS, (T1)= 1.7 set ...................327
Figure C.3 Comparison of the target CS and the selected ground motions:
(a) geometric mean spectral values; (b) logarithmic standard deviation of
spectral values .............................................................................................330
Figure C.4 Response spectra of ground motions in the CS, (1) = 1.8 set ..................330

xxviii
Chapter 1

Introduction

1.1 Motivation and background


Life safety and collapse prevention have always been primary goals of earthquake
engineering. Until recently this protection against collapse was not explicitly quantified
but instead was assumed to be sufficient for structures designed to standards specified by
building codes. Advances in computational power and the development of models that
can more accurately reproduce the behavior of structural components through failure
have made collapse risk assessment possible. With the advent of performance-based
earthquake engineering (PBEE), which considers uncertainties in the seismic hazard and
structural response and seeks to engineer structures so that they achieve a desired level of
performance in terms of expected monetary losses, downtime and casualties (e.g.,
Krawinkler and Miranda 2004; Deierlein 2004), collapse risk assessment has become
increasingly important. In fact, it is a necessary part of the PBEE methodology as
collapse contributes to the expected economic losses, downtime and casualties resulting
from a seismic event.

Though collapse risk assessment is possible it is not widely performed, partially due to
the computational demands. Even when considering a single seismic event or ground
motion intensity level, the response of a given structure can vary widely due to inherent
variability from one ground motion to the next. This means that analyzing the structure
using many ground motions is often required to accurately predict the response.
Quantifying the structural response to different seismic events or ground motion intensity
levels can require hundreds of dynamic, nonlinear response analyses. There are also
many uncertainties involved in the risk assessment process (e.g., uncertainty in seismic
hazard or structural properties), and considering the effect of these uncertainties can
further add to the computational effort.

1
Chapter 1: Introduction 2

In recent years some design guidelines have started providing information regarding the
expected performance of a structure as it relates to collapse. For example, ASCE 7-10
(ASCE 2010, Table C1.3.1b) provides the anticipated maximum probability of collapse
conditioned on a given level of ground motion intensity for structures designed according
to these guidelines. ASCE 7-10 also targets a 1% probability of collapse in 50 years when
determining the spectral values used in design. Recent studies such as those by Haselton
and Deierlein (2007, Chapter 6) have shown, however, that even buildings designed for
the same site using the same code, materials, and structural system can have strikingly
different collapse risks due to variances in design decisions (e.g., structural layout or the
distribution of strength and stiffness over the height). This means that without some type
of analysis the collapse risk of a particular structure remains largely unknown.

1.2 Objectives and scope


This dissertation focuses on seismic collapse risk assessment of buildings and evaluates
the effects of ground motion intensity measure (IM) selection, which is used to quantify
the seismic hazard and predict the structural response, and computational approach on the
computed risk. The main objectives of this research are:
(1) To evaluate different metrics for quantifying the collapse risk and to describe the
advantages and disadvantages of each metric.
(2) To examine and quantify the uncertainty in the collapse fragility, which describes
the probability of collapse conditioned on ground motion intensity, and the mean
annual frequency of collapse (c) due to the number of ground motions considered
in structural response analysis.
(3) To develop efficient and reliable procedures for estimating the collapse risk.
(4) To explain why certain IMs are able to predict structural response better (i.e.,
more efficiently) than others.
(5) To evaluate the performance of different IMs with respect to predicting the
collapse risk. Particular focus is given to the efficiency and sufficiency of the IM
and how these properties affect the computed collapse risk.
Chapter 1: Introduction 3

(6) To study spectral acceleration averaged over a period range (Saavg) as an IM and
provide recommendations on the period ranges that maximize the efficiency of
the IM for different structures.

Though there are different types of structural collapse, this dissertation focuses on
collapse due to seismic events and in particular the sidesway mode of collapse in which
the lateral displacement of a story or number of stories due to P- effects and component
deterioration causes dynamic instability. This is opposed to other types of collapse such
as loss of vertical carrying capacity or progressive collapse, which can occur when
structural elements are effectively removed due to blast-type loading.

As previously mentioned there are many sources of uncertainty in the collapse risk
assessment process, including uncertainty in the seismic hazard, structural response, and
structural modeling (e.g., material strengths or choice of hysteretic behavior). This
dissertation primarily focuses on uncertainty in the collapse fragility and c estimates
arising from the number of ground motions used in structural analysis, which is
sometimes known as statistical uncertainty. The effects of record-to-record variability,
which are related to the former, are also examined; however, the effects of modeling
uncertainty when estimating the seismic hazard and defining the structural model are not
considered (i.e., expected/mean seismic hazard curves are used and only a single model is
used for a given structure).

The discussion of collapse risk metrics and the uncertainty in the collapse fragility due to
the number of ground motions considered as well as the procedures for estimating the
collapse risk are general to any type of structure. The discussion focused on the IMs can
be structure-specific, however. In particular, the period range that maximizes the
efficiency of Saavg can vary between structures. This dissertation focuses on low- to mid-
rise (i.e., one to twenty story), medium period range (i.e., fundamental period between
0.5 s and 3.2 s) structures whose lateral force-resisting systems are either moment-
resisting frames or shear walls that fail in flexure.
Chapter 1: Introduction 4

1.3 Organization
This dissertation addresses topics related to seismic collapse risk assessment, focusing on
how the IM selection and computational approach affect the efficiency and reliability of
the computed collapse risk estimate.

Chapter 2 discusses different methods for performing a collapse risk assessment. It


begins with an in-depth review of the current state of knowledge in analytical collapse
simulation and describes two methods for evaluating collapse risk: a fully probabilistic
method and an IM-based method. Different collapse risk metrics are presented and
compared with respect to their advantages and disadvantages. Finally, the chapter
presents an overview of selected simplified methods used in assessing the collapse risk of
structures.

Chapter 3 presents a complete collapse risk assessment of a case study structure. Details
of this case study including site characteristics, ground motion selection, structural
properties, and modeling assumptions are discussed in depth as this case study is used
throughout the dissertation. In addition to evaluating of the collapse risk of the structure,
the deaggregation of the collapse risk by intensity level is presented, and the use of this
type of analysis is described. The effect of the shape of the seismic hazard curve on the
deaggregation is investigated using example sites from the western and central/eastern
Unites States.

Chapter 4 focuses on the statistical uncertainty in the collapse fragility and the mean
annual frequency of collapse (c) stemming from the number of ground motions used to
predict the structural response. Theoretical confidence intervals on the probability of
collapse conditioned on ground motion intensity are presented as well as confidence
intervals on the collapse fragility and c for the case study described in Chapter 3 that are
created from bootstrap (repeated, random sampling) techniques. A novel method for
estimating the collapse risk based on two intensity levels is also proposed. The premise of
this method is that, for an equivalent amount of computational effort, a better (less
uncertain) estimate of the collapse risk can be obtained by considering more ground
Chapter 1: Introduction 5

motions at fewer intensity levels compared to using many intensity levels but fewer
ground motions at each level. The proposed method is illustrated using a case study.

Chapter 5 investigates the relationship between spectral shape and structural response,
focusing on identifying the features of a ground motion that make it particularly
damaging to a certain structure. This is approached through a number of methods
including the investigation of individual terms of the equation of motion, the response of
a system to a half-sine acceleration pulse, the peak displacement response of a system to
a ground motion, and the investigation of the spectral shape of damaging records.
Damaging records are identified for the case study presented in Chapter 3. A metric for
quantifying the spectral shape of a record is identified, and it is shown to be strongly
correlated with the intensity to which a record must be scaled in order to produce collapse
in a given structure. Methods for incorporating this metric in collapse risk assessment
analyses are discussed.

Chapter 6 evaluates the performance of different IMs in predicting the collapse risk of a
structure. Properties of the IM such as efficiency, sufficiency, and the ability to compute
hazard information are discussed, and these properties are evaluated for an IM that
measures the average spectral acceleration over a period range (Saavg). Nearly 700
structures are used to evaluate the performance of Saavg and to compare its properties to
those of more traditional IMs. Finally, the collapse risk predictions computed by different
IMs are evaluated using a case study structure with different ground motion sets.

Chapter 7 builds on the analysis of Saavg presented in Chapter 6 by identifying the period
ranges that maximize the efficiency of this IM. Nearly 700 structures are used to
determine the influence of structural properties on the optimum period range. The
number of periods used to compute Saavg and the spacing of periods within the period
range are also considered.

Chapter 8 presents the summary and major conclusions of this dissertation. It also
describes the limitations of the present study and provides suggestions for future work.
Chapter 2

Assessing Collapse Risk

2.1 Overview
This chapter explores various methods of quantifying the collapse risk of a structure due
to seismic excitations. It begins with a review of analytical research on structural collapse
focused primarily on the sidesway mode of collapse in which the lateral displacement of
a story or number of stories due to P- effects and component deterioration causes
dynamic instability. This is opposed to other types of collapse such as loss of vertical
carrying capacity or progressive collapse, which can occur when structural elements are
effectively removed due to blast-type loading. The review of analytical research on
collapse builds on existing literature reviews provided by Ibarra and Krawinkler (2005,
Section 2.2) and Villaverde (2007). The chapter then then discusses various approaches
for evaluating the collapse risk. Different metrics used to quantify collapse risk are also
presented, and the advantages and disadvantages of each metric are discussed.

2.2 Previous analytical studies on structural collapse


Some of the earliest analytical research on structural collapse focused on collapse in
single-degree-of-freedom (SDOF) systems due to geometric nonlinearities (P- effects)
in which the effect of gravity load P acting on the displaced structure is considered. One
of the first studies was performed by Jennings and Husid (1968) who investigated the
time required to produce collapse in SDOFs. Though the systems had positive post-yield
stiffness, incorporation of P- effects effectively created a negative post-yield stiffness
that permitted collapse. They found that the time to collapse was strongly related to the
height of the structure, the yield strength, and the post-yield stiffness. Sun et al. (1973)
studied elasto-plastic SDOFs and found that the collapse of a structure is directly related
to its yield displacement and stability coefficient (the ratio of the gravity load P to the
product of lateral stiffness and the height of the structure). Takizawa and Jennings (1980)
examined the collapse intensity of SDOFs and studied the effects of ground motion

6
Chapter 2: Assessing Collapse Risk 7

duration. This was one of the earliest studies to investigate collapse under the combined
effects of P- and material degradation. They used SDOFs with a trilinear force-
displacement backbone curve to represent the pre-cracked, cracked/pre-yield, and post-
yield behavior of reinforced concrete (RC) frame structures. The post-yield region was
characterized by constant strength; however, they also considered a quadrilinear force-
displacement backbone curve in which the fourth region had negative stiffness to
represent the loss in strength due to spalling and crushing of concrete at large
deformations. The authors found that including the strength-loss region significantly
reduced the collapse intensity.

Other studies that investigated the collapse of non-deteriorating SDOFs (i.e., SDOFs
without cyclic deterioration of strength or stiffness) due to P- effects include those by
Bernal (1992), Kanvinde (2003), Vian and Bruneau (2003), Miranda and Akkar (2003),
and Adam and Jger (2012a), although in these studies the SDOF is intended to serve as a
proxy to a multiple-degree-of-freedom (MDOF) system. Bernal (1992) used elasto-plastic
SDOFs and SDOFs with stiffness-degrading hysteretic behavior to develop expressions
for the minimum base shear required to prevent collapse under a given ground motion as
a function of the structural properties (the period and stability coefficient) and ground
motion properties (peak ground velocity or displacement and duration). Kanvinde (2003)
tested 19 steel frame SDOF systems to collapse on a shake table and showed that
analytical models could accurately and reliably predict the results. He used concentrated
plasticity elements with the Giufr-Menegotto-Pinto hysteretic model, which uses a
nonlinear hardening law. His experimental tests complemented those by Vian and
Bruneau (2003), who tested 15 steel frame SDOF systems to collapse on a shake table.
Vian and Bruneau found that the stability coefficient was strongly related to the structural
performance. Miranda and Akkar (2003) computed probabilistic estimates of collapse for
SDOFs with bilinear force-displacement backbones and negative post-yield stiffness by
using a relative intensity measure defined as the ratio of the lateral strength required for
the system to remain elastic to the minimum lateral strength required to avoid dynamic
instability under a given record. They developed expressions for the mean collapse
strength ratio of a system and quantified the record-to-record variability as a function of
Chapter 2: Assessing Collapse Risk 8

its period and the ratio of post-yield stiffness to elastic stiffness. More recently, Adam
and Jger (2012a) used SDOFs with bilinear force-displacement backbones to develop
expressions for median collapse capacity spectra as a function of period, effective post-
yield stiffness, damping, and hysteretic behavior.

Early studies on collapse using non-deteriorating MDOF systems include those by


Chopra et al. (1973), Osteraas and Krawinkler (1990), Ger et al. (1993), Challa and Hall
(1994), and Gupta and Krawinkler (1999). Chopra et al. (1973) studied the collapse of the
Olive View Medical Center main building during the 1971 San Fernando earthquake.
Using a two-dimensional (2D) model, the authors represented the moment-curvature
relationships of concrete members using a bilinear hysteretic model with a yield stiffness
equal to 3% of the initial elastic stiffness. They were able to reproduce some of the
observed damage when analyzing the model using a simulated ground motion; however,
the computed displacements were smaller than those imposed by the earthquake. The
authors attributed the discrepancies to use of a simulated ground motion in place of the
actual ground motion at the site, which is unknown, and the redistribution of forces
caused by failure of elements very early in the earthquake that were not included in the
model. They concluded that the severity of the ground motion and the termination of
shear walls below the second floor, which caused a large change in strength and stiffness,
were the major reasons for the severe damage. Both Osteraas and Krawinkler (1990) and
Ger et al. (1993) studied the collapse of a 22-story structure in the Pino Suarez complex
during in the 1985 Mexico City earthquake. Osteraas and Krawinkler (1990) analyzed a
2D model using a portion of the recorded ground acceleration and duplicated the
observed failure mode of the building. The model included inelastic springs that
represented the buckling of truss girder webs as well as hysteric models representing the
buckling of braces and the 4th-story columns. They studied the effects of different
modeling assumptions (e.g., stiff vs. flexible, different post-buckling residual strengths,
and the strength of girder-to-column connections) and found that the inelastic response
was highly dependent on the post-buckling behavior of the columns. Ger et al. (1993)
used a three-dimensional structural model and subjected it to the recorded ground
motions from that earthquake using two orthogonal horizontal acceleration components
Chapter 2: Assessing Collapse Risk 9

as well as the vertical component. The hysteretic models capture strength loss in the
beams and the effects of local buckling on the axial and flexural response of the columns
as well as the axial response of brace elements. Only when including the effects of both
local buckling in the columns and ductile failures in the beams do the authors obtain
results consistent with field observations of identical buildings near the collapsed
structure. Challa and Hall (1994) examined the performance of a code-conforming, 20-
story steel moment-resisting frame (MRF) and used fiber sections to model the beams
and columns in a 2D model. The hysteretic model consisted of strain-shifting backbone
curves that included some strength degradation after reaching the ultimate strength, and
the cyclic behavior was governed by elliptical unloading and reloading curves. They
found that very severe ground motions were required to collapse the structure; however,
they noted that structural deterioration, which was not included, could be a very
important factor affecting the response of the structure. Gupta and Krawinkler (1999,
Chapter 4) also evaluated performance of a 20-story steel MRF designed according to the
1994 Uniform Building Code (UBC 1994) using a 2D model. Nonlinear behavior was
represented with concentrated plasticity elements, and all inelastic elements had a bilinear
force-displacement backbone with 3% strain-hardening. The authors found that the
structure was potentially susceptible to collapse due to P- effects and that the structural
response was sensitive to modeling assumptions (e.g., inclusion of panel zones and the
effects of slabs and interior gravity framing). They also stressed the importance of using
hysteretic models that capture strength and stiffness degradation to better assess the
structural response at large deformations.

Song and Ellingwood (1999b; 1999a) studied the seismic reliability of special steel
moment frames with welded connections, with a particular focus on the effects of
connection fractures and modeling uncertainty associated with the hysteric behavior and
other structural properties including the yield strength of beams and columns and level of
damping. They modeled a 4-story building in California designed in 1985 and used a
bilinear hysteric model for columns, both bilinear and degrading hysteretic models for the
beams, and a degrading hysteretic model for the connections. Using a set of simulated
and recorded ground motions, they found that the probability of incipient collapse
Chapter 2: Assessing Collapse Risk 10

(defined as exceeding an interstory drift ratio or roof drift ratio of 5%) estimates were
significantly affected by whether a bilinear or degrading hysteric model was assumed for
the connection behavior. The authors found that the onset of connection cracking and the
reduced capacity of the connection were the two parameters of the connection model that
had the greatest influence on structural response but that the uncertainty in hysteresis was
small compared to the uncertainty in ground motions, with the recorded ground motions
producing more variation in structural response than the simulated motions. The authors
found that the annual probability of incipient collapse was on the order of 10-4.

Sivaselvan and Reinhorn (2002; 2006) and Sivaselvan et al. (2009) have applied
Lagrangian or mixed Lagrangian formulations to analyze the collapse behavior of
structures. Using a case study of a portal frame subjected to dynamic loads, Sivaselvan
and Reinhorn (2006) demonstrated that the Lagrangian formulation can produce
significantly different vertical displacements than the more traditional displacement-
based solutions. The authors also noted that the Lagrangian formulation is particularly
powerful when determining the redistribution of forces and momentum in a structure
when individual elements collapse. Sivaselvan et al. (2009) demonstrated that the
collapse behavior of a case study structure with elasto-plastic members subjected to an
earthquake could be captured using a relatively large time step (0.02 s) based on the fact
that the behavior converged to what was predicted using time step ten times smaller
(0.002 s). Lavan et al. (2009) extended the mixed Lagrangian formulation to account for
strength degradation and fracture, and they demonstrated that the method was stable and
that the results of fracture simulations converged even when using a relatively large time
step.

Lee and Foutch (2002) evaluated the global drift capacity of structures under a set of
ground motions using incremental dynamic analysis (IDA) (Vamvatsikos and Cornell
2002). IDA involved repeating nonlinear response history analyses using increasing
levels of ground motion intensity until collapse occurs. The hysteretic model governing
bending in the beams was based on a trilinear force-displacement backbone including a
negative stiffness region. The authors defined collapse as: (1) dynamic instability when
Chapter 2: Assessing Collapse Risk 11

the drift increases without bounds; or (2) the point at which the slope of the IDA curve
(the plot of ground motion intensity versus maximum drift) becomes less than 20% of the
slope from elastic analysis; or (3) the point when the maximum drift reaches 0.10, beyond
which confidence in the results of the analytical model are questionable. They analyze
structures of 3, 9, and 20 stories, but only the 20-story structure experiences collapse
before the drift limit of 0.10. They find that the median global drift capacity of the 20-
story structure is 0.08 with a logarithmic standard deviation of 0.26 when using W24
columns. Jalayer (2003, Chapter 4) also uses the IDA approach to compute the global
drift capacity of a 7-story reinforced concrete frame structure in a probabilistic format.
The nonlinear behavior is represented by rotational and translational (shear) springs that
capture degradation in strength and stiffness. She defines collapse as the point at which
the slope of the IDA curve becomes less than 20% of the elastic slope. The global drift
capacity is combined with a drift hazard curve based on seismic hazard data at the site to
compute the mean annual frequency of exceeding the global drift capacity.

Medina and Krawinkler (2003, Chapter 9) also used IDA to evaluate the collapse
performance of a structure under a set of records. The structure is an 18-story frame that
does not include any cyclic strength or stiffness deterioration. Unlike the other studies,
they define the collapse capacity of the structure in terms of the ground motion intensity
(not the drift level) at which collapse occurs. They measure the ground motion intensity
by Sa(T1), the spectral acceleration at the structures first-mode period. They describe the
collapse capacities via a collapse fragility curve, which is a lognormal cumulative
probability distribution fit to the collapse intensities of all records in the ground motion
set and describes the probability of collapse conditioned on the ground motion intensity.
The collapse fragility curve concept is also discussed by Ibarra et al. (2002). The collapse
fragility data is combined with seismic hazard data describing the mean annual frequency
of exceeding the ground motion intensity to give the mean annual frequency of collapse
(c). This method of computing the collapse risk is outlined more formally by Ibarra and
Krawinkler (2005, Chapter 7). It should be noted that the former reference was originally
published in the same year as the Medina and Krawinkler study as (Ibarra 2003).
Chapter 2: Assessing Collapse Risk 12

Ibarra and Krawinkler made significant contributions to the collapse risk assessment of
structural systems. They developed a hysteretic model that captured basic modes of
cyclic strength and stiffness deterioration and was calibrated based on experimental
component tests (Ibarra and Krawinkler 2005; Ibarra et al. 2005, Chapter 3; Ibarra et al.
2005). The model was based on a quadrilinear backbone curve including elastic, strain-
hardening, post-capping (strength loss), and residual strength regions. Deterioration
modes include basic strength and post-capping deterioration as well as unloading and
accelerated reloading stiffness deterioration. This hysteretic model is referred to as the
Ibarra-Medina-Krawinkler (IMK) model. Ibarra and Krawinkler (2005, Chapter 6; 2011)
also examined the effect of both record-to-record (RTR) variability and modeling
uncertainty (i.e., uncertainty in parameters of the hysteretic model) on the collapse risk.
They found that incorporating modeling uncertainty could increase c by 50% to 100%
compared to only considering RTR variability and that the capping displacement and
post-capping stiffness had very significant contributions to modeling uncertainty.

An approximate approach for including epistemic uncertainty (modeling uncertainty) in


collapse risk assessment is discussed by Zareian and Krawinkler (2007a), which involves
using an estimate of the dispersion on median collapse intensity due to epistemic
uncertainty to either shift (decrease) the median collapse intensity or to inflate the
dispersion on the median collapse intensity. Liel et al. (2009) also studied the effect of
modeling uncertainty on collapse risk assessment using both first-order, second-moment
(FOSM) methods and more complicated Monte Carlo sampling with a response surface.
They propose an approximate method for incorporating modeling uncertainty based on
balancing the relative computational efficiency of FOSM with the more accurate but
computationally demanding response surface method. Dolek (2009) proposed using
Monte Carlo simulation with Latin hypercube sampling (LHS) to reduce the
computational demands in estimating the uncertainty in collapse predictions due to
modeling uncertainty.

Vamvatsikos and Fragiadakis (2010) assessed the seismic performance of a case study
structure when incorporating modeling uncertainty using three approaches: Monte Carlo
Chapter 2: Assessing Collapse Risk 13

simulation with LHS, point estimates, and FOSM. They found that point-estimate and
FOSM methods could allow accurate performance estimates compared to the more
computationally demanding Monte Carlo simulation with LHS up to maximum interstory
drift ratios of 5%, but they observed that the median IDA curves computed by former two
methods tended to oscillate at higher drift ratios and predict smaller collapse intensities
than the LHS method. They also found that the performance estimates could be very
sensitive to the uncertainty in the ultimate rotational ductility of the beam elements,
which were the only elements that experienced nonlinear behavior in their particular case
study. They observed that while the median collapse intensity found using the mean
(deterministic) values of model parameters was larger than the median collapse intensity
found when considering modeling uncertainty, the difference was only 10%. Finally, they
observed that the overall uncertainty in the response (which considers both RTR
variability and modeling uncertainty) found directly from the Monte Carlo results was
generally well approximated by the square-root-of-the-sum-of-the-squares approach for
combining RTR variability and modeling uncertainty, especially when the maximum
interstory drift ratio exceeded 10%.

Haselton and Deierlein (2007) performed collapse risk assessments for modern, code-
conforming (ASCE 2005; ICC 2005; ACI 2005) RC MRFs designed for California, while
Liel and Deierlein (2008) did the same for older, non-ductile RC frames that represented
typical structures built between 1950 and 1975 in California. The authors used a portfolio
of buildings with a variety of structural configurations, number of stories, design
decisions, etc. and assessed how these differences affected the collapse risk. Haselton and
Deierlein found that the average c was 3x10-4 for modern, code-conforming structures
while Liel and Deierlein found that the average c was 40 times greater for older, non-
ductile structures. Both studies simulated nonlinear behavior via concentrated plasticity
elements that used the IMK deterioration model. Haselton and Deierlein (2007, Chapter
4) calibrated parameters of the deterioration model based on a series of 255 experimental
tests on RC columns and developed predictive equations of the model parameters based
on regression of these results.
Chapter 2: Assessing Collapse Risk 14

Ellingwood et al. (2007) studied the seismic fragilities of low- to mid-rise steel and RC
MRFs representative of structures in the central and eastern United States, which
typically are designed primarily for gravity and wind resistance. They modeled four
structures: two steel and two RC. For the steel MRFs nonlinear behavior was represented
by distributed plasticity elements with a bilinear hysteretic response and 3% strain
hardening for steel members and, in one of the steel frames, a hysteretic model for the
partially restrained (PR) connections characterized by a decreasing strength region
following capping. The RC MRF models used fiber sections for the beams and columns
and a beam-column joint model that represented the effects of joint shear and bond-slip.
Using simulated ground motions, they concluded that while steel MRFs with PR
connections may not collapse at substantial ground motion intensities, the RC MRFs may
not survive the design-basis ground motion without experiencing significant damage or
collapse. Li and Ellingwood (2007) also examined the seismic fragilities of single-story,
wooden shear wall structures representative of residential construction in the United
States. They compared the nonlinear responses of three wooden shear walls: one solid
wall, one with a door-sized and a window-sized opening, and the final with a garage-
sized opening. The hysteretic behavior of the models incorporated the pinching behavior
observed in experimental tests. Using a suite of ground motions the authors found that
inelastic response of each structure was virtually indistinguishable from the linear
response, even up to drift ratios of 5% (the largest observed value), which was
unexpected because the equal displacement rule typically does not work well for short
period structures like the 0.16 s to 0.32 s structures examined here. They found that the
average probability of exceeding the collapse prevention state corresponding to a 3% drift
ratio was approximately 0.01% for the solid shear wall and nearly three times this value
for the shear walls with openings.

Christovasilis et al. (2009) studied the seismic collapse fragility of woodframe structures,
particularly the effects of construction quality, excitation direction, and wall finishes. The
authors modeled a two-story townhouse and a three-story apartment building and
included nonlinear shear springs with a pinching hysteresis model that captured strength
and stiffness degradation. The models were analyzed using IDA, with collapse defined as
Chapter 2: Assessing Collapse Risk 15

exceeding a 7% interstory drift ratio. The authors found that using bi-directional ground
motions decreased the median collapse intensity by approximately 20% compared to
analyzing the structure with a ground motion in only one direction. They also observed
that the construction quality significantly affected the collapse fragility and that including
wall finishes in the structural model generally, but not always, decreased the probability
of collapse conditioned on ground motion intensity.

The National Institute of Standards and Technology (NIST) (2010) released a report
evaluating the FEMA P695 (ATC 2009a) methodology for quantifying building
performance factors and assessing the collapse risk of a variety of structural systems,
which is intended for use in determining appropriate design criteria for seismic force-
resisting systems. The original FEMA report considered special and ordinary RC MRFs
and light-framed walls with wood panel sheathing. The NIST report evaluated seven
additional structural systems, including special and ordinary shear walls constructed of
RC or reinforced masonry, special steel MRFs and concentrically braced frames, and
buckling-restrained braced frames. An acceptable level of collapse performance in the
NIST study was based on the adjusted collapse margin ratio (ACMR), which is calculated
by multiplying a spectral shape factor by the ratio of the median collapse intensity (i.e.,
the ground motion intensity at which the probability of collapse is 50%) to the ground
motion intensity at the maximum considered earthquake (MCE) level. An acceptable
level of collapse performance is essentially achieved when the average probability of
collapse given the MCE is less than or equal to 10% for a given performance group (i.e.,
a set of archetype structures with a common structural system and behavior) and less than
or equal to 20% for any individual archetype within the group. The NIST report
concluded that most systems designed according to ASCE 7-05 (ASCE 2005) would have
an acceptable level of collapse performance, although notable exceptions were short-
period RC or reinforced masonry shear walls and short-period steel braced frames. The
NIST report concluded that using period-dependent seismic performance factors may be
necessary.
Chapter 2: Assessing Collapse Risk 16

Krishnan and Muto (2012) studied the collapse mechanisms of two 18-story steel MRF
structures using 3D models. They subjected the structures to idealized ground motions
represented by a series of identical rectangular acceleration pulses, and they repeated the
analyses using different pulse amplitudes, periods, and number of cycles. The idealized
ground motions were intended to represent near-fault, pulse-like ground motions. They
found that of all the possible collapse mechanisms, only a limited number prevailed and
that these mechanisms were likely to occur in stories where where the strength of the
structural elements is low enough, but the driving mass of the overriding floors is
sufficiently large. The authors also demonstrate that the principle of virtual work can be
used with plastic analysis to predict the characteristic collapse mechanisms.

Wang et al. (2012) demonstrated the viability of hybrid numerical and experimental
simulation by reproducing the shake table results of a full scale, 4-story steel MRF
structure tested to collapse. During the shake table test the first-story columns
experienced large levels of nonlinearity leading to the formation of a first-story collapse
mechanism, so in the hybrid test the first story was physically tested while the rest of the
structure was modeled numerically. The two physical specimen substructures, which
were tested at different locations, consisted of a 1.5-bay by 1.5-story specimen and a 0.5-
bay by 1.5-story specimen. The hybrid simulation was able to capture the first-story
collapse mechanism observed during the shake table test; however, the comparison of
hysteretic behavior at the first story showed significant differences between the shake
table and hybrid tests. The authors attributed the differences to a number of factors
including loading in only one direction for the hybrid test versus tri-directional excitation
in the shake table test, boundary assumptions in the hybrid test, presence of non-
structural components and a concrete slab in the shake table test, and different material
strengths in the shake table and hybrid specimens. Overall, they concluded that the hybrid
test adequately captured the distribution of forces in the frame and produced a similar
response and collapse mechanism to what was observed in the shake table test.

Lignos and Krawinkler (2012a, Chapter 2) modified the IMK deterioration model (Ibarra
et al. 2005) to include the ability to simulate a complete loss of strength and redefined
Chapter 2: Assessing Collapse Risk 17

some of the models input parameters to make them less sensitive to modeling
uncertainties. They also added the ability to define different backbone curves and
deterioration rates in the positive and negative loading directions in order to simulate the
effect of composite action on the collapse capacity of steel MRFs (Lignos et al. 2011a).
Lignos and Krawinkler (2012b) provide statistical information related to the correlation
of various modeling parameters that may be used to more reliably assess the effect of
modeling uncertainties on the collapse capacity of steel and RC buildings. Lignos and
Krawinkler (2011; 2012a, Chapter 3) calibrated the input parameters of the modified
IMK deterioration model based on experimental tests for steel W-sections and square
tubular columns as well as reinforced concrete beams and developed predictive equations
of the model parameters based on regression of these results. For steel W-sections,
separate equations are developed for beams with and without reduced beam sections
(RBS). Lignos and Krawinkler (2012a, Chapter 7) and Lignos et al. (2011b) also
analytically predicted and reproduced the results of shake table tests of a scaled frame
structure up through collapse, demonstrating that component deterioration significantly
affects the collapse risk of frame structures and that collapse can be predicted fairly well
using relatively simple models. When comparing blind analysis predictions to the
experimental results of a shake table collapse test of a full-scale 4-story steel MRF,
Lignos et al. (2013) found that including both P- effects and the deterioration of steel
components were critical for reliably predicting collapse. The authors also found that
Rayleigh damping yielded better response predictions than stiffness-proportional
damping and that for predicting the sidesway collapse mechanism of the regular plan
building they studied, a 3D model of the structure had no clear advantage over a 2D
model.

A number of recent studies have focused on the assessing the collapse potential of steel
concentrically braced frames (CBFs), including those by Stoakes (2012), Hsiao et al.
(2013), Li and Fahnestock (2013). Building on previous work by Hines et al. (2009),
Stoakes (2012) showed buildings with reserve lateral force-resisting systems (LFRS),
achieved primarily from the flexural strength in the beam-column connections, showed
improved collapse resistance for CBF systems designed in a moderate seismic zone. He
Chapter 2: Assessing Collapse Risk 18

found that the strength of the LFRS was more significant than the stiffness in improving
the collapse resistance. Hsiao et al. (2013) compared the performance of a 3-story, a 9-
story, and a 20-story CBF building designed for Seattle, WA. They found that at the 2/50
hazard level (i.e., the hazard level with a 2% chance of exceedance in 50 years) including
the effects of the gravity frame had a significant effect on the response of the 3-story
building but not the 9-story or 20-story buildings. In a case study of the 3-story and 20-
story buildings, they found that the median collapse intensity (the ground motion
intensity where half of the ground motions cause collapse) was, on average, 20% higher
when ground motions were scaled to the same Sa(T1) versus scaling based on the record
set intensity, which is the median Sa(T1) of the unscaled records (i.e., when the intensity
was measured using Sa(T1) versus the median Sa(T1) of the record set.) After evaluating
the structures using the FEMA P695 (ATC 2009a) methodology, the authors recommend
that the response modification coefficient (commonly called the R factor) for special
CBFs specified by ASCE 7-10 (ASCE 2010) should be reduced from R = 6 to R = 3 for
low-rise CBF buildings in order for their margin of safety against collapse to be more
consistent with mid-and high-rise CBF buildings. Li and Fahnestock (2013) examined the
collapse resistance at the maximum considered event (MCE) level defined in ASCE 7-10
(ASCE 2010) of low-ductility CBFs, which are representative of CBFs currently
constructed in moderate seismic regions and older CBFs in high seismic regions. They
used SDOF models that included P- effects and reserve capacity to model the resistance
of the structure after the brittle failure of a brace or brace connection. They found that
increasing the reserve capacity was more effective at improving the collapse resistance
than decreasing the R factor and that the ductility of the reserve system is typically more
important than the global drift capacity.

Karamanci (2013) and Lignos and Karamanci (2013) have also performed research
related to collapse risk assessment of CBFs. They found that realistic simulation of the
nonlinear response of a CBF requires that the stiffness-proportional term used in
Rayleigh damping must be computed based on the current stiffness rather than the initial
stiffness because the damping forces produced when using the initial stiffness are not
realistic after flexural buckling and subsequent brace fractures (Karamanci 2013; Chapter
Chapter 2: Assessing Collapse Risk 19

6). Furthermore, in a case study of a 2-story special CBF they found that using the initial
stiffness was unconservative as the median collapse intensity of this structure was
reduced by over 40% when using the current stiffness. The authors also found that the
RTR variability in the median collapse intensity was approximately 0.6, which is
significantly greater than the typical value of approximately 0.4 for steel MRFs (Ibarra
and Krawinkler 2005; Zareian and Krawinkler 2007b). The authors also found that the
collapse risk, as measured by the mean annual frequency of collapse (c), of special CBFs
is slightly larger than that of code-compliant special steel and RC MRFs obtained in other
studies. They believe this is because damage tends to concentrate in stories where the
steel braces have fractured, making CBFs more susceptible to local story collapse
mechanisms. Finally, Karamanci (2013, Chapter 5) and Lignos and Karamanci (2013)
have developed predictive equations for modeling the cyclic buckling and fracture of
HSS, round HSS and W-shaped steel braces that were calibrated from hundreds of
experimental tests on steel braces.

It should be noted that numerous studies have examined how the spectral shape of ground
motions affects collapse risk assessment. These studies are discussed in Chapter 5.

2.3 Methods for evaluating collapse risk


Various methods exist for evaluating the collapse risk of a structure. Conceptually the
collapse risk could be computed by observing a given structure over a period of
thousands of years and counting the number of times it collapses (assuming it is
immediately repaired or rebuilt each time it is damaged or collapses). This information
could be used to calculate the collapse risk in terms of an average rate or probability of
collapse over a certain period of time. Although conceptually correct, this method is, of
course, impractical. Another straightforward but also impractical method would be to
create an analytical model that perfectly represents the structure, subject it to ground
motions recorded at the site over a period of thousands of years, and count the number of
collapses. Because the oldest recorded strong ground motion accelerogram is from 1933
(Trifunac 2009), this method is also not feasible. It is, however, feasible to use ground
Chapter 2: Assessing Collapse Risk 20

motions recorded at other sites or simulated ground motions provided they appropriately
represent the types of ground motions expected to occur at the site.

Because the ground motions a structure will experience in its lifetime are unknown, the
collapse risk of a structure can be estimated by evaluating its response to various ground
motions and weighting the responses by how frequently each type of ground motion is
expected to occur at the site. This approach for calculating the collapse risk is described
mathematically by Bazzurro et al. (1998) as

PC vi P(C | event )i (2.1)


i

where PC is the annual probability of collapse, i is the mean annual rate of events on
fault i greater than a given magnitude (e.g., Mw > 5), and P(C | event)i is the probability
of collapse given an event on fault i. The latter quantity can be estimated as the fraction
of records causing collapse, where the records represent an event on the fault and are
selected or simulated based on the joint probability distribution of magnitude and
distance for the fault.

The former approach is used by Wen (1995) and Jalayer et al. (2004), for example, to
compute the probability of exceeding a given value of inter-story drift ratio (IDR) over a
given time period. The authors of both studies select or simulate records that are
consistent with the magnitudes and distances of the seismic sources surrounding their
sites. This fully probabilistic method can require an extremely large number of ground
motions to capture the range of potential future events, although as noted by Jalayer et al.
(2004) smart sampling techniques (e.g., LHS) can significantly reduce the number of
ground motions required.

An alternative approach involves characterizing ground motions by a scalar intensity


measure (IM) such as Sa(T1), which is currently a commonly used IM. The structural
response is evaluated using ground motions with a range of intensity levels. Different sets
Chapter 2: Assessing Collapse Risk 21

of ground motions can be used at different intensity levels, or the same set of ground
motions can be scaled to different intensity levels. The collapse risk is estimated by
combining the structural response (in this case, the probability of collapse) conditioned
on the intensity of the ground motion with seismic hazard data describing how frequently
each intensity level is exceeded at the site. This approach for calculating the collapse risk
is described mathematically as

c =

0
P(C | im ) d IM (im ) (2.2)

where c is the mean annual frequency of collapse, P(C | im) is the probability that the
structure will collapse when subjected to an earthquake with ground motion intensity
level im, and IM is the mean annual frequency of exceedance of the ground motion
intensity measure im. The advantage of this approach is that |dIM(im)| is readily available
for most sites as IM is the main output of a probabilistic seismic hazard analysis (PSHA).

An implicit assumption in this approach is a Markovian dependence in which it is


assumed that structural collapse depends solely on the intensity level of the ground
motion, as measured by an intensity parameter, and not on other parameters that may
influence the ground intensity measure. This property is called sufficiency and it means
that if, for example, Sa(T1) is used as the IM, the response of the structure or probability
of collapse at a given intensity level is not influenced by the magnitude, distance, or
duration of the event the produced that ground motion. This is advantageous because it
means that ground motions from a variety of events can be used for an intensity level of
interest. The assumption of sufficiency may not be entirely correct as, for example,
Tothong and Cornell (2006) observed a small dependence between the peak displacement
of inelastic SDOFs and earthquake magnitude when records were scaled to the same
value of Sa(T1); however, many other studies have concluded that Sa(T1) is sufficient
with respect to magnitude and distance for predicting maximum IDRs (e.g., Shome et al.
1998; Bradley et al. 2010a). Other studies have shown that Sa(T1) is not sufficient with
Chapter 2: Assessing Collapse Risk 22

respect to a spectral shape parameter called 1 (e.g., Baker and Cornell 2006a; Haselton et
al. 2011), meaning that the intensity level at which the ground motion causes the structure
to collapse depends on the value of . Baker and Cornell (2005a, Chapter 5) also note that
Sa(T1) is not sufficient with respect to the period of the velocity pulse in near-fault, pulse-
like ground motions. In such cases where the IM is not sufficient with respect to a ground
motion parameter, some modification needs to be made so that the risk assessment results
are valid. Modifications including using a vector IM and careful record selection are
discussed by Baker and Cornell (2005a).

Another issue related to the sufficiency of the IM is whether the same structural response
estimates at a given intensity level are the same when using unscaled/as-recorded ground
motions versus ground motions that are scaled to the intensity level. Luco and Bazzurro
(2007) examined this particular issue with respect to the median drift estimates obtained
for nonlinear SDOF systems and one MDOF system. Their objective was to determine
whether, for a given narrow magnitude and distance (Mw-R) bin and a given level of
Sa(T1), using records scaled to the target level of Sa(T1) produces a bias in the median
drift response compared to using unscaled records with the target magnitude, distance,
and unscaled Sa(T1) values. They sorted records into six Mw-R bins and compared the
results obtained from a given target Mw-R bin to the results obtained using records from a
different Mw-R bin that were scaled to match the Sa(T1) value of the target bin. The
authors found that scaled records could introduce significant bias in the drift response,
which they attributed to differences in the median elastic spectral shape of the different
bins. They noted that the bias could be significantly reduced by selecting records to
match the target spectral shape. They also looked at scaling records within a given bin
and found that, on average, records scaled up to a given intensity level produced larger
responses than records that did not need to be scaled (i.e., that were recorded with the
given intensity level). Baker (2007a) demonstrated that for records scaled to the same
Sa(T1) value, the scale factor did not influence the maximum interstory drift ratio of a
seven-story RC MRF structure (i.e., there was no bias) when the records were selected

1
The spectral shape parameter measures the number of standard deviations between the spectral ordinate
at a given period of a particular ground motion and the geometric mean value predicted by a ground motion
prediction equation. More information about this parameter is provided, e.g., by Baker and Cornell (2006).
Chapter 2: Assessing Collapse Risk 23

based on spectral shape (either via the parameter or by matching a mean spectral shape).
He found that a bias due to scale factor did exist, however, when records were selected
arbitrarily or from a given Mw-R bin. Bradley et al. (2010a) also observed that no bias to
due scale factor in the maximum interstory drift results of a 10-story RC MRF structure
when using records scaled to a common Sa(T1) value that were selected based on spectral
shape.

When comparing the fully probabilistic method of collapse assessment and the IM-based
approach, Baker and Cornell (2005a, Chapter 1) note:

The concerns with simulated ground motions, the added computational expense
of performing many analyses and the complications involved when the ground
motion hazard and structural response cannot be treated independently mean that
the intensity measure approach still has many advantages as a method for
assessing seismic risk to structures.

For these and other reasons, the scalar IM-based approach is currently widely used to
assess the seismic risk of structures and is at the core of the performance-based
earthquake engineering methodology used by the Pacific Earthquake Engineering
Research Center (Deierlein 2004). The scalar IM-based approach will be used throughout
this work unless noted otherwise.

2.4 Evaluation of different collapse risk metrics


This section describes different metrics used to quantify collapse risk and discusses the
pros and cons of each.

2.4.1 Probability of collapse at a specific ground motion intensity,


P(C|IM)
One way to measure the collapse risk of structure is by computing its probability of
collapse conditioned on a specific ground motion intensity (P(C|IM)), such as the
intensity associated with a 2% probability of exceedance in 50 years (the 2/50 hazard).
For example, Haselton and Deierlein (2007) reported the probability of collapse
Chapter 2: Assessing Collapse Risk 24

conditioned on the intensity associated with 2/50 hazard (P(C|2/50)) as one of the many
metrics they used to describe the collapse risk of RC frame structures. Positive attributes
of this metric include its ability to be computed relatively easily compared to other
metrics discussed later (i.e., it involves computing the probability of collapse at a single,
pre-defined intensity level) and the fact that it combines information about the structure
and the seismic hazard at the site (assuming the ground motion intensity is tied to the
hazard level).

The main disadvantage of this metric is that the collapse risk is only measured at a single
level of ground motion intensity when, due to uncertainties in the structural capacity,
multiple intensity levels contribute to the collapse risk. A collapse fragility curve is a
cumulative distribution function that describes how the probability of collapse increases
with increasing ground motion intensity. Measuring the collapse risk by P(C|IM) is
equivalent to evaluating the collapse risk at only one point on the collapse fragility curve
(only at one level of ground motion intensity), and therefore it does not incorporate
information about changes in the probability of collapse with changes in the ground
motion intensity.

Consider a situation in which two different structures with the same fundamental period
of vibration are located at the same site. Since both structures have the same period and
are located at the same site their seismic hazard curves would be identical even if using a
structure-dependent intensity measure such as Sa(T1). Figure 2.1 shows hypothetical
collapse fragility curves for the two structures, Structure A and Structure B. Even though
the seismic hazard is identical, whether Structure A is safer (i.e., has a lower collapse risk
as measured by the P(C|IM) value) than Structure B depends on the IM level one
considers. For example, Structure A is more likely to collapse (i.e., it has a larger
P(C|IM) value) than Structure B when the IM level is greater than 1, but the opposite is
true for IM levels less than 1. The relative difference in the collapse risk of the two
structures as measured by P(C|IM) also depends on the IM level as, for example, the
structures have comparable collapse risks around an IM level of 1, but at an IM level of
0.5 Structure B is approximately eight times more likely to collapse than Structure A.
Chapter 2: Assessing Collapse Risk 25

Figure 2.1: Hypothetical collapse fragility curves for two structures located at the same
site.

Consider another situation in which identical structures (i.e., with the same collapse
fragility function) are located at different sites with the same 2/50 hazard. This is
illustrated in Figure 2.2, which shows the hypothetical collapse fragility curve in Figure
2.2(a) and the seismic hazard curves for the two sites, Site A and Site B, in Figure 2.2(b).
The collapse risk as measured by P(C|2/50) will be the same for both cases, but due to
differences in the seismic hazards at the sites (e.g., different slopes in the seismic hazard
curves at the 2/50 level), the collapse risks will be different when a different hazard level
is used. For example, at more frequent hazard levels (i.e., larger values of IM) than the
2/50 level the structure at Site A will have a greater probability of collapse than the
structure at Site B because the ground motion intensity is larger at Site A. The opposite is
true when considering hazard levels less frequent than the 2/50. The degree to which the
relative collapse risks vary with the chosen hazard level depends on differences in the
seismic hazard curves at the two sites. If the shapes of the hazard curves are similar, the
relative collapse risks may be relatively stable; however, if the shapes of the hazard
curves are very different, the relative collapse risks can vary significantly with the chosen
hazard level. This is illustrated in Section 3.5 for six sites located throughout the United
States.
Chapter 2: Assessing Collapse Risk 26

Figure 2.2: Hypothetical structure located at two different sites: (a) collapse fragility
curve; (b) seismic hazard curves that cross at the 2/50 hazard level.

Given that P(C|IM) yields variable collapse risk results in the special situations
previously discussed, it is clear that this metric is not the most appropriate for comparing
the collapse risk of different structures located at different sites.

On a related note, some design provisions in the United States used to base the design of
a structure on the spectral ordinate associated with a specific hazard level. As discussed
by Luco et al. (2007), this led to a variable collapse risk depending on the location of the
structure. For this reason, more recent design provisions such as ASCE 7-10 (ASCE
2010, Commentary C1) now use a risk-targeted spectral ordinate (rather than a hazard-
targeted spectral ordinate) that is aimed at achieving a more uniform collapse risk among
structures.
Chapter 2: Assessing Collapse Risk 27

2.4.2 Median collapse intensity


Another collapse risk metric is the median collapse intensity, which is the ground motion
intensity at which the structure has a 50% probability of collapse. As this metric is not
associated with the sites seismic hazard, it is not appropriate for comparing the collapse
risk of structures in different locations (i.e., with different seismic hazards) or for
assessing the relative collapse risk of different structures at a given site whose
fundamental periods of vibration differ significantly. However, it can be used to
approximately compare the collapse risks of different structures located at the same site,
provided the same IM is used (e.g., if spectral acceleration is used it needs to be
evaluated at the same period for both structures so that the IM hazard is the same) and the
dispersion on the median collapse intensity is approximately the same for both structures.
The latter condition is necessary because the dispersion in a structures capacity is related
to the range of intensities that contribute to the collapse risk (e.g., if the dispersion on the
collapse intensity were very small, the collapse risk would be dominated by the median
collapse intensity, but if significant dispersion existed the risk would be distributed
among many intensities).

Figure 2.1 shows how the dispersion affects the collapse fragility curves of two structures
with the same median collapse intensity. Structure B, which has more dispersion than
Structure A, has a larger probability of collapse than Structure A at intensities less than
the median but a smaller probability of collapse than Structure A at intensities greater
than the median. Whether Structure A is safer than Structure B once all intensities are
considered depends on the hazard associated with each intensity, which, loosely
speaking, determines whether the majority of the collapse risk comes from intensities less
than or greater than the median. More information about identifying the contribution of
each intensity to the collapse risk is provided in Section 2.4.4.
Chapter 2: Assessing Collapse Risk 28

2.4.3 Collapse margin ratio (CMR)


The collapse margin ratio (CMR) is defined as the ratio of the median collapse intensity
to another intensity that is typically associated with a specific hazard level. The 2/50
hazard level or the Maximum Considered Event (MCE), which typically corresponds to
the 2/50 hazard, are commonly used (e.g., ATC 2009a; NIST 2010; Liel et al. 2011).
Note that both FEMA P695 (ATC 2009a) and NIST (2010) use an adjusted CMR, which
multiplies the CMR by a spectral shape factor that depends on the first-mode period of
the structure, a period-based ductility determined via a pushover analysis, and the seismic
design category. The CMR can be viewed as a factor of safety against a structure having
a 50% probability of collapse given the MCE or 2/50 hazard. In a sense the CMR
combines the previous two metrics (except it uses the intensity associated with the MCE
or 2/50 hazard level rather than the probability of collapse at that intensity), and it
addresses some of their shortcomings. However, the CMR also has some of the same
problems as the previous metrics because it only considers two intensities that contribute
to the collapse risk (the intensities associated with the median collapse and the MCE or
2/50 hazard) and only incorporates hazard information from a single point on the seismic
hazard curve (e.g., it ignores the slope of the hazard curve at that point and the seismic
hazard at all other intensities).

It is possible to have structures with the same CMRs but different collapse risks as
measured by c, which considers all intensities that contribute to the collapse risk, even
for special cases in which the structures are located at the same site (Case A) or are
identical (Case B). An example of Case A is illustrated in Figure 2.1, which shows the
collapse fragility curves of two structures with the same median collapse intensity but
different levels of dispersion. It is assumed that the same IM (i.e., hazard curve) is used
for both structures, meaning that the CMRs are equal because the IM hazards are equal.
Different collapse risks as measured by c would arise because of different levels of
dispersion on the median collapse intensity, which relates to how significantly each
intensity contributes to the collapse risk as described in Sections 2.4.2 and 2.4.4. Case B
could arise because of differences in the seismic hazard at intensities other than the MCE
Chapter 2: Assessing Collapse Risk 29

or 2/50 hazard level as described in Section 2.4.1 and illustrated by Figure 2.2(b), which
shows two seismic hazard curves that cross at the 2/50 level.

The CMR can be a useful metric because, like P(C|IM) when the IM is associated with a
specific hazard level, it provides information about the site and the structure. It is harder
to compute than P(C|IM) because the intensity level associated with the median collapse
is unknown a-priori; however, the CMR generally provides a better estimate of the
relative collapse risk than P(C|IM) because it considers two intensities rather than one.

2.4.4 Mean annual frequency of collapse2 (c)


As previously mentioned, the mean annual frequency of collapse (c) considers all
intensities that contribute to the collapse risk, and it also accounts for differences in the
seismic hazard and the structural capacity and associated uncertainty. For these reasons, it
can be used to directly compare the collapse risk of different structures located at
different sites and is the best metric for evaluating the collapse risk.

Computation of c (which is described by Equation 2.2) requires more information and


effort than the other collapse risk metrics. Unlike the CMR, c requires the full
distribution of the structures collapse capacity as a function of IM (i.e., knowledge of the
full fragility curve) and of the full seismic hazard curve with respect to intensity (not just
values at particular intensities) based on a number of simplifying assumptions. Some
researchers (e.g., Shome and Cornell 1999; Jalayer 2003; Zareian and Krawinkler 2007a;
Bradley and Dhakal 2008) have proposed approximate closed-form solutions for
evaluating c or the mean annual frequency of exceeding a limit state, which has a similar
computational format. The closed-form solutions are based on a lognormal distribution of
the collapse fragility curve and a seismic hazard curve that is linear in log-log domain. As
discussed by Aslani and Miranda (2005b) or more recently by Bradley and Dhakal (2008)
and Vamvatsikos (2012), errors in the closed-form solution of c with respect to the value

2
The material in this section and the following section is based on
Eads, L., Miranda, E., Krawinkler, H., and Lignos, D. G. (2013). "An efficient method for estimating
the collapse risk of structures in seismic regions." Earthquake Engineering & Structural Dynamics,
42(1), 25-41.
Chapter 2: Assessing Collapse Risk 30

computed via numerical integration are mostly due to the representation of the hazard
curve as linear in the log-log domain and to a lesser extent the dispersion on the median
collapse intensity. These studies recommend evaluating c via numerical integration as it
is straightforward, represents a negligible computational effort with respect to evaluating
the structural response at a given intensity level, and is not subject to the errors
introduced by the closed-form expression.

The following paragraphs explain the process of obtaining c through numerical


integration and how to conduct a deaggregation of c by intensity to identify each
intensitys contribution to the collapse risk. As previously stated, two pieces of
information are needed to calculate c: the seismic hazard curve, which gives the mean
annual frequency of exceeding different ground motion intensities at the site, and the
structures collapse fragility curve, which describes the structures probability of collapse
conditioned on the intensity of the ground motion. Computation of c is described by
Equation 2.2 and is expanded to demonstrate how to determine each intensitys
contribution to the collapse risk. When multiplying and dividing the right-hand side of
Equation 2.2 by d(im), calculation of c can be rewritten as


d IM (im )
c =

0
P(C | im )
d (im )
d (im ) (2.3)

where dIM(im)/d(im) is the slope of the seismic hazard curve at the site. In general, there
is no closed-form solution to the integral in Equation 2.3, and therefore this integral is
typically solved using numerical integration. This is achieved by computing the product
of the probability of collapse conditioned on IM and the slope of the seismic hazard curve
at discrete IMs, multiplying by the increment in IM (im), and finally adding the results
from all IMs. This process is represented in Equation 2.4 and is illustrated graphically in
Figure 2.3:
Chapter 2: Assessing Collapse Risk 31


d IM (imi )
c =

i =1
P(C | imi )
d (im )
im (2.4)

Figure 2.3: Illustration of c deaggregation process: (a) collapse fragility curve;


(b) numerical derivative of seismic hazard curve; (c) c deaggregation.

A deaggregation of c by intensity provides a powerful tool for identifying the


contribution of different levels of ground motion intensity to the total collapse risk. This
is analogous to the deaggregation by magnitude, distance, and epsilon () used in PSHA
to determine the seismic sources primarily contributing to the hazard.

A point on the c deaggregation curve is obtained by plotting the product of the


probability of collapse at a given ground motion intensity and the slope of the seismic
hazard curve as a function of intensity. As indicated by Equation 2.4, the contribution of
a given (small) range of ground motion intensities to c is obtained by multiplying the
probability of collapse at the intensity corresponding to the midpoint of the intensity
range by the slope of the seismic hazard curve at that intensity and by the width of the
Chapter 2: Assessing Collapse Risk 32

ground motion intensity range im. As illustrated schematically in Figure 2.3(c), the area
under the deaggregation curve is equal to c. Ground motions intensities with higher
ordinates in the deaggregation curve indicate higher contributions to c.

Figure 2.4 shows that c deaggregation analysis allows the identification of the range of
intensities that primarily contribute to c. For example one may determine the intensities
im1 and im2 that bracket a range of ground motion intensities contributing to, say, 75%
of c.

Figure 2.4: Illustration of identifying the contribution to c from a range of intensities:


(a) collapse fragility curve; (b) c deaggregation.

Figure 2.3 also illustrates that ground motion intensities associated with high probabilities
of collapse do not necessarily have large contributions to c because they occur less
frequently than intensities in the lower half of the collapse fragility curve. Therefore, as
will be shown in Sections 3.5 and 3.6, the largest contribution to c typically comes from
ground motion intensities in the lower half of the collapse fragility curve because even
though these intensities have relatively small probabilities of collapse, the slope of the
Chapter 2: Assessing Collapse Risk 33

seismic hazard curve is typically much larger for these intensities than for those in the
upper half of the collapse fragility curve.

2.4.5 Probability of collapse over a given time period


The final collapse risk metric discussed is the probability of collapse over a given time
period. This metric is essentially a different representation of c that is easier to
communicate with project stakeholders. If the occurrence of collapses in time is assumed
to follow a Poisson process (a common assumption in earthquake engineering), the
probability of one collapse over t years can be computed as follows

Pc (in t years ) = 1 exp( c t ) (2.5)

Because c is a small value for typical buildings, the annual probability of collapse is
approximately equal to c, that is

Pc (in 1 year ) c (2.6)

and the probability of collapse over t years is approximately equal to

t
1
Pc (in t years ) 1 1 (2.7)
c

Given that the design life of typical structures is 50 years, many owners and stakeholders
might be interested in their structures probability of collapse over a 50-year period. For a
structure to have less than a 1% probability of collapse in 50 years (the target for
structures designed according to ASCE 7-10 (ASCE 2010), c must be less than or equal
to 2.0x10-4 according to Equation 2.5.
Chapter 2: Assessing Collapse Risk 34

2.5 Simplified collapse risk assessment


Generally, estimating the collapse risk of a structure involves trade-offs between
computational effort and accuracy. A number of simplified, approximate methods exist
for assessing the collapse risk of structures that reduce the computational effort involved.
Most of these methods focus on obtaining an estimate of the median collapse intensity,
and many also provide estimates of the dispersion for characterizing the full distribution
of the collapse fragility curve. Some of these approximate methods were mentioned in
Section 2.2, while others are described in this section. It should be noted that the
approximate methods discussed in this chapter are not an exhaustive list. Approximate
methods for assessing the collapse risk of structures can generally be divided into two
categories: those that require the user to perform nonlinear response history analyses
(RHAs) and those that do not. Methods that do not require RHAs are discussed first.

Vamvatsikos and Cornell (2006), FEMA P440A (ATC 2009b), Han et al. (2010), Shafei
et al. (2011), Fajfar and Dolek (2012), and Adam and Jger (2012b), for example, have
all developed methods for estimating the median collapse intensity that are based on a
pushover analysis of the MDOF structure and do not require performing RHAs.
Vamvatsikos and Cornell (2006) developed the static pushover to IDA (SPO2IDA)
software tool which yields estimates of the median collapse intensity as well as the 16th
and 84th percentiles based on results from a static pushover of the structure. The user
inputs the static pushover results as an idealized backbone curve (up to a quadrilinear
curve is permitted), and SPO2IDA returns the collapse intensity predictions, which are
based on regressions of IDA results for SDOF systems with various backbone shapes
using 30 ground motions. Equation 4-4 in FEMA P440A (ATC 2009b) provides an
equation for estimating the median collapse intensity based on a pushover curve of the
structure. The equation includes terms for systems with and without stiffness degradation
and was calibrated based on the IDA results of SDOF systems. Han et al. (2010) provide
empirical equations for the median collapse intensity and the dispersion as a function of
the period, damping, and properties of a first-mode-based SDOF that is constructed via a
modal pushover analysis (Chopra and Goel 2002) of the MDOF structure. The equations
Chapter 2: Assessing Collapse Risk 35

are based on regression of the IDA results of 4800 SDOF systems and 240 ground
motions.

Shafei et al. (2011) also provide empirical equations for the median collapse intensity that
use properties of an idealized pushover curve of the structure and the type of structure
(MRF or shear wall) as inputs. The pushover curve is based on the lateral load pattern
used by ASCE 7-05 (ASCE 2005). The equations are based on regression of the IDA
results of MDOF models representing generic MRF structures and generic shear wall
structures with a variety of characteristics including height, lateral stiffness, yield base
shear coefficient, cyclic deterioration parameters of individual components, etc. under a
set of 40 ground motions. The equations also permit consideration of how the spectral
shape parameter affects the median collapse intensity. Estimates of dispersion are also
provided based on the structural type.

The final method discussed was proposed by Adam and Jger (2012b), which provides
expressions for 16th, 50th (median), and 84th collapse intensity percentiles of regular,
MRF structures where collapse is dominated by P- effects. The inputs for the
expressions are based on properties of an equivalent bilinear SDOF system derived from
a pushover of the MDOF system and consider different types of hysteretic behavior. The
expressions are based on regression of IDAs on SDOF systems defined by a bilinear
force-displacement backbone that did not consider cyclic deterioration.

Simplified collapse risk assessment methods that require performing RHAs include those
by Han and Chopra (2006), Azarbakht and Dolek (2007; 2010), Eads et al. (2013), and
Baker (2013). These methods, compared to those described above that do not require
RHAs, require more computational effort; however, they are typically more accurate
because they capture the dynamic response of MDOF structures. Han and Chopra (2006)
proposed an approximate method for replicating IDA curves that is based on performing
RHAs on inelastic SDOF systems. Each SDOF system represents a different mode of the
structure, and its properties are determined from a modal pushover analysis of the MDOF
structure. The SDOF responses are combined using modal combination rules to
Chapter 2: Assessing Collapse Risk 36

approximate the response of the MDOF structure. The authors found that using two or
three modal systems was generally appropriate for approximating the MDOF response.

Azarbakht and Dolek (2007; 2010) developed a method for estimating the 16th, 50th
(median), and 84th fractile IDA curves for first-mode-dominated structures. The method
involves first running IDAs on an SDOF system that represents the MDOF structure. The
results are used to develop a precedence list of ground motions, which is a subset of the
ground motions that best replicate the fractile curves and is determined via optimization
using a genetic algorithm or other approach. Progressive IDAs are then performed on
the MDOF structure in which ground motions are selected according to the precedence
list. After performing IDAs with two ground motions for a desired fractile IDA curve
(e.g., the 50th fractile curve), the median IDA curve for the desired fractile is calculated
and is updated after analyzing each additional ground motion. The analysis stops when
the change in the median IDA curve for the desired fractile is less than some tolerance. In
a case study, the authors found that this method predicted the collapse intensity with a
10% error using only 15 ground motions compared to the intensity calculated from the
original set of 98 ground motions, which represented an 80% reduction in computational
effort.

The methods proposed by Eads et al. (2013) and more recently by Baker (2013) focus on
estimating points on the collapse fragility curve at a limited number of intensity levels
(typically two or three) to achieve a good estimate of c. These intensities are identified
through a c deaggregation using an initial estimate of the collapse fragility curve. RHAs
are then performed on the MDOF structure at the selected intensities. Estimates of
uncertainty related to the number of ground motions used in RHA are also provided for
select case studies. Baker also provides recommendations on how to select the intensity
levels if the objective is to obtain a good estimate of the collapse fragility curve (not c).
The method proposed by Eads et al. is discussed in detail in Section 4.5.
Chapter 2: Assessing Collapse Risk 37

2.6 Summary
This chapter reviewed how analytical research on structural collapse has progressed over
time. It then described two methods for evaluating collapse risk: a fully probabilistic
method and an IM-based method. The IM-based method is currently widely used to due
to the reduced computational effort, among other reasons, and is the method used in this
document. The chapter continued by describing different collapse risk metrics and
comparing the pros and cons of each. It was seen that c (or the related metric of the
probability of collapse over a given time period) is the best metric for describing the
collapse risk because it combines information about the structure and the site and
accounts for differences in the seismic hazard and the collapse fragility at all intensities.
The chapter concluded with an overview of selected simplified methods for assessing the
collapse risk of structures.
Chapter 3

Collapse Risk Assessment Case Study:


4-Story Steel Frame Structure

3.1 Overview
This chapter provides details about a collapse risk assessment case study of a four-story
steel frame structure located in the Los Angeles, California area whose results are used
extensively throughout this document. Information about the site, seismic hazard, ground
motion selection, structure and modeling, and simulation results are discussed. Much of
the material in this chapter is based on a paper by Eads et al. (2013), although notable
differences include use of: (1) a different hazard curve due to the availability of updated
hazard data since the paper was submitted and the ability to directly generate data for the
site class of interest; and (2) a different collapse fragility curve due to changes in the
definition of damping for the structural analysis model.

3.2 Site and seismic hazard analysis


The Bulk Mail Center (33.996 N, 118.162 W) located south of downtown Los Angeles
in Bell, California is chosen as the site for this case study because it has been studied
extensively in previous studies. This location represents a typical site in urban California
with high seismicity that is not dominated by unusually strong near-fault effects even
though it is within 20 km of seven faults. Detailed information about the characteristics
of this site is provided by Haselton et al. (2008, Chapters 3 and 4), but some of the more
basic attributes of the site discussed by Haselton et al. are repeated here. Figure 3.1 shows
the faults contributing to the hazard at the site for T = 1 s. The seismic hazard is
controlled by strike-slip and reverse or reverse/oblique faults. The average shear wave
velocity in the upper 30 m of soil at the site is Vs-30 = 285 m/s, which classifies the site as
a NEHRP Site Class D (BSSC 2003).

38
Chapter 3: Collapse Risk Assessment Case Study 39

Figure 3.1: Map of Bulk Mail site (yellow star) and surrounding faults. Faults controlling
the site hazard at T = 1 s are shown in blue (strike-slip) and red (reverse or
reverse/oblique) while other faults contributing to the hazard are shown in white. Thin
yellow lines denote freeways. (from Haselton et al. 2008).

A probabilistic seismic hazard deaggregation obtained from the 2008 Interactive


Deaggregations web site by the United States Geological Survey (USGS) (USGS 2012b)
is shown in Figure 3.2. This figure shows the magnitudes (Mw), distances (R), and
epsilons () contributing to the spectral acceleration at T = 1 s (Sa(T = 1 s)) with 2%
probability of exceedance in 50 years. The intensity level with this mean rate of
exceedance is 1.05 g. A period of T = 1 s was used because it is the period closest to the
first-mode period of the case study structure (T1 = 1.33 s) for which deaggregation data is
available. As seen in Figure 3.2 the seismic hazard is dominated by sources close to the
site, and the mean values of (Mw, R, ) are (6.79, 8.9 km, 1.1).
Chapter 3: Collapse Risk Assessment Case Study 40

Figure 3.2: Probabilistic seismic hazard deaggregation for Bulk Mail Site for Sa(T = 1 s)
with a 2% probability of exceedance in 50 years (USGS 2012b).

The seismic hazard curve for Sa(T1) is obtained from the Hazard Curve Application from
the USGS (USGS 2012a) using data from the 2008 National Seismic Hazard Mapping
Project (Petersen et al. 2008). Because seismic hazard data is only available for select
periods, the hazard data from the Hazard Curve Application for Sa(T = 1 s) and Sa(T = 2
s) for Site Class D conditions was interpolated to give the hazard data at Sa(T1 = 1.33 s),
the first-mode period of the case study structure. Because neither the reported hazard
levels nor the reported intensity levels were the same for the 1-s and 2-s data, each data
set was fit with a polynomial curve in log-log space to facilitate interpolation. A 4th-order
polynomial was judged to sufficiently represent the data for this purpose as it captured
the shape of the hazard curve and the discrete data points provided by the Hazard Curve
Application. Points on the 1.33-s hazard curve were found by linear interpolation in log-
log space of the 1-s and 2-s curves at discrete hazard levels. To facilitate collapse risk
assessment calculations, the discrete seismic hazard data points were fit with a
continuous function. The function shows how the (slope of the) seismic hazard curve
Chapter 3: Collapse Risk Assessment Case Study 41

changes between the discrete points and facilitates using small intensity increments when
calculating c via numerical integration. The points on the 1.33-s hazard curve were fit
with a 4th-order polynomial in the log-log domain, which was again judged to sufficiently
represent the data. The resulting seismic hazard curve for Sa(T1) is presented in Figure
3.3, which shows the mean annual frequency () of exceeding a given value of Sa(T1) at
the site. For reference, the intensity corresponding to a 2% probability in 50 years is
Sa(T1) = 0.83 g. The 1-s and 2-s hazard curves used in interpolation are also shown in
Figure 3.3 for reference.

Figure 3.3: Seismic hazard curve for the Bulk Mail Center site.

3.3 Ground motion selection


The objective of ground motion selection is to choose ground motions representative of
the motions expected at the site and that can be used to compute c. The motions selected
should be consistent in terms of expected frequency content, duration, spectral shape, and
other relevant features such as the presence of pulse-like motions if directivity effects are
anticipated. Additional information about record selection is provided by Baker and
Cornell (2006a). A large sample of damaging motions (e.g., those in the lowest 10% of
the collapse fragility curve) is particularly important for Chapter 5, which focuses on
identifying characteristics of damaging ground motions and distinguishing them from
more benign ground motions, and using many records minimizes the likelihood of
Chapter 3: Collapse Risk Assessment Case Study 42

conclusions being sensitive to record selection. In order to accomplish these objectives,


the following ground motion selection criteria were considered.

Ground motions were selected from the PEER Next Generation Attenuation (NGA)
database (Power et al. 2008) based on magnitude, distance, focal mechanism, and site
class. It is important to note that no additional criteria such as were included in order to
facilitate studies of spectral shape and structural response in Chapter 5. The set is
comprised of all NGA records, excluding those from dam abutments, that are within a
specified range of magnitude (Mw between 6.93 and 7.62), Joyner-Boore distance
(between 0 and 27 km), NEHRP site class (C or D), and are from strike-slip, reverse, or
reverse-oblique faults. The ground motion set consists of 137 acceleration records (each
with two horizontal components) from 11 different events. The set includes records with
a variety of characteristics, including pulse-like (records affected by forward directivity)
and non-pulse-like records as seven faults are located within 20 km of the site. Additional
details about the records in this set including NGA identification numbers, station names,
and event data are provided in Appendix A. This ground motion set is herein and
henceforth referred to as MRCD 137 as it was selected based on magnitude (M),
distance (R), NEHRP Site Class (C and D), and 137 records comprise the set. The
distance range used in record selection approximates the sites main hazard contributors
from the deaggregation in Figure 3.2. The site is NEHRP Site Class D, and Site Class C
records are included to expand the ground motion set. The selected fault mechanisms are
consistent with the mechanisms of the faults dominating the hazard at the site (Haselton
et al. 2008). The magnitude range of this set (Mw between 6.93 and 7.62), however, is not
consistent with the magnitudes dominating the site hazard in Figure 3.2. The magnitude
range was adjusted from what was expected at the site in favor of including the numerous
recordings from the Mw = 7.62 Chi-Chi, Taiwan earthquake of 1999, which was
beneficial for the purposes of Chapter 4 and Chapter 5. Implications of the magnitude
range adjustment on the collapse risk estimate are discussed in Chapter 6.
Chapter 3: Collapse Risk Assessment Case Study 43

3.4 Description of structure and modeling


This section provides details about the structure including structural system, geometry,
and design parameters as well as a description of the analytical model of the structure
used for response history analysis.

3.4.1 General description of structure


The structure used in this case study is a four-story office building whose lateral force-
resisting system consists of steel special moment-resisting frames with reduced beam
sections (RBS). Full documentation of the design is provided by Lignos and Krawinkler
(2012a, Chapter 5), but a summary of the major details is provided here. The structural
system is designed in accordance with the 2003 International Building Code (IBC) (ICC
2003) and the 2005 AISC seismic provisions (AISC 2005b; a). The design spectral
acceleration ordinates SDS and SD1 are 1.0 g and 0.6 g, respectively. A plan view of the
structure is shown in Figure 3.4(a), and an elevation view of the moment-resisting frames
along gridlines 1 and 4, which are the focus of this case study, is shown in Figure 3.4(b).
The design base shear coefficient in the direction parallel to the numbered gridlines
(Figure 3.4a) is V/W = 0.082. The first three modal periods of the building along gridlines
1 and 4 are T1 = 1.33 s, T2 = 0.43 s, and T3 = 0.22 s based on the analytical model
described in the following paragraph.
Chapter 3: Collapse Risk Assessment Case Study 44

(a)

(b)
Figure 3.4: Case-study structure: (a) plan view; (b) elevation view along gridline 4 (from
Lignos and Krawinkler 2012a).

3.4.2 General description of model


The moment-resisting frame on gridline 4 between gridlines B and D (which is identical
to the moment-resisting frame on gridline 1) (Figure 3.4) is modeled using an in-house
version of the Open System for Earthquake Engineering Simulation (OpenSees)
(McKenna 1997; 2009). The analysis program uses OpenSees Version 2.1.1 as a base
with add-ons implemented by Lignos including a hysteretic model and stiffness modifiers
for elastic elements, both of which are described later. A concentrated plasticity concept
is used with the frame members modeled as elastic elements with nonlinear rotational
springs at their ends as discussed by Lignos and Krawinkler (2012a) and Lignos et al.
(2011b). Concentrating inelastic behavior at the ends of beam-column elements at a
single point is a simplification of an elements actual behavior, but it is a common
assumption that has been made by many researchers modeling the collapse of structures
Chapter 3: Collapse Risk Assessment Case Study 45

(e.g., Haselton et al. 2008; Zareian and Krawinkler 2009; Lignos and Krawinkler 2012a),
which, if adequately implemented, does not introduce significant errors in the global
dynamic response of the structure (Lignos and Krawinkler 2012a). The nonlinear
rotational springs are zero-length elements (point springs) located where the plastic
hinges are expected to form (i.e., at the center of the RBS locations for the beams and at
the edges of the panel zones for the columns). The model is based on centerline
dimensions, and the frame columns are fixed at the base. In order to reduce the
computational effort when evaluating the response at multiple ground motion intensity
levels when subjected to a large number of ground motions, panel zones were not
explicitly modeled.

3.4.3 Mass and loading


The seismically effective weights of the second, third, fourth, and fifth/roof floors are
1070 kips, 1050 kips, 1050 kips, and 1200 kips, respectively (Lignos and Krawinkler
2012a, Table 5.3), for a total weight of 4370 kips. This weight includes the dead load and
a portion of the live load. As there are two identical frames in the direction of interest, the
modeled frame is assigned half of the seismically effective weight (i.e., it carries 2185
kips). The seismic mass is applied at the beam-column joints in each floor and distributed
equally among each joint in the floor. Gravity loads based on the seismic weight are
applied as point loads at the beam-column joints in each floor and are distributed as
follows: the frame columns receive gravity load based on tributary area, and the
remaining gravity load in the floor (i.e., gravity loads carried gravity-only columns) is
lumped on a leaning column. This leaning column simulates P- effects and is connected
to the frame by axially-rigid truss elements. Any lateral resistance provided by the gravity
system is neglected as the leaning column is pinned at the base and between floors.

3.4.4 Damping
Rayleigh damping is used with 2% of critical damping assigned to the first and third
modes. Zareian and Medina (2010) showed that Rayleigh damping can lead to unrealistic
damping forces for concentrated-plasticity systems and can cause significant
underestimation of the collapse intensity. To overcome this problem, they propose
Chapter 3: Collapse Risk Assessment Case Study 46

applying stiffness-proportional damping only to the elastic elements, which involves


modifying the damping coefficient and the stiffness matrix of the elastic beam element to
ensure that the total damping work is not affected. The approach to Rayleigh damping
described by Zareian and Medina (2010) is adopted. The nonlinear rotational springs are
made ten times stiffer than the rotational stiffness of the beam element following the
approaches of Medina and Krawinkler (2003, Appendix A) and Ibarra and Krawinkler
(2005, Appendix B), which concentrates the elastic deformation in the elastic beam
element and the inelastic deformations in the nonlinear rotational springs. It should be
noted that stiffness-proportional damping is only assigned to the elastic frame elements,
not the leaning column or to the truss members linking the frame to the leaning column.

3.4.5 Hysteretic behavior


The hysteretic behavior of the nonlinear rotational springs is governed by a bilinear
hysteretic response with a modified version of the Ibarra-Medina-Krawinkler
deterioration model (Ibarra et al. 2005; Lignos and Krawinkler 2011; Lignos and
Krawinkler 2012a, Chapter 2). The model utilizes a multi-linear force-displacement
capacity curve and cyclic deterioration parameters that govern stiffness and strength
deterioration as shown in Figure 3.5. The force-displacement capacity curve includes an
elastic region, a post-yield/pre-capping region between yielding and maximum strength
(capping strength), and a post-capping region. The post-capping region can include a
residual strength region where the strength of the element is constant and drops to zero at
a specified ultimate rotational capacity. The residual strength region represents the
behavior of steel components when local buckles stabilize before ductile tearing, and a
residual strength ratio of = 0.4 is recommended based on limited experimental results
(Lignos and Krawinkler 2011). The cyclic deterioration modes include basic strength
deterioration between yield and the maximum strength (the capping strength), post-
capping strength deterioration, unloading stiffness deterioration, and accelerated
reloading stiffness deterioration, although the latter mode does not apply to the bilinear
hysteretic response used in the case study.
Chapter 3: Collapse Risk Assessment Case Study 47

Figure 3.5: Modified Ibarra-Medina-Krawinkler deterioration model and definition of


parameters (from Lignos et al. 2011b).

Empirical relationships for defining parameter values of this model were developed by
Lignos and Krawinkler (2011) via multivariate nonlinear regression analyses of
experimental results from over 300 steel specimens. The parameters used for the
nonlinear rotational springs are provided in Table 3.1 and are based on the properties
provided by Lignos and Krawinkler (2012a, Table 6.1). The yield moments are calculated
based on the expected plastic moment, where the expected plastic moment at the RBS
location is used for the beams. Changes in moment capacity due to axial-moment
interaction effects are not considered as the analytical software does not have this
capability. As explained by Lignos and Krawinkler (2011), the parameter defines the
cyclic deterioration modes and is accompanied by an exponent c that controls the rate of
cyclic deterioration and another parameter D that permits different deterioration rates in
the positive and negative loading directions. Both the c and D parameters are defined as
1, the default value, for the case study. The value of the latter parameter implies that the
effect of the slab is neglected, which is the case as no composite action between the beam
and slab is considered (e.g., there is no increase in the flexural stiffness or expected
moment capacity of the beam due to composite action). It has been shown that including
the contribution of the slab may affect the collapse intensity as the increase in beam
bending strength can change the collapse mechanism from a distributed mechanism to
Chapter 3: Collapse Risk Assessment Case Study 48

more local story mechanism (Lignos et al. 2011a; Elkady and Lignos 2013); however,
slab effects are neglected for this case study.

Table 3.1: Modeling parameters for nonlinear rotational springs used in case study*.
Story or Ke (kip- p pc
Element Section My (lb-in)
Floor in/rad) (rad) (rad)
3.5 - 4 Column W24x76 11000 32838 0.025 0.35 1.5
2 - 3.5 Column W24x131 20350 65819 0.025 0.30 1.5
1 Column W24x131 20350 46226 0.025 0.30 1.5
4, 5 (roof) Beam W21x93 10938 12974 0.025 0.19 1.9
2, 3 Beam W27x102 7930 23251 0.020 0.16 1.5
*These parameters are defined in Figure 3.5. All springs use = 0.4; Mc/My = 1.05; and u = 0.4 rad.

The type of modeling used here, specifically the concentrated plasticity concept with
modified Rayleigh damping and use of the modified Ibarra-Medina-Krawinkler hysteretic
model, has been shown to successfully replicate the collapse behavior of both scaled
(Lignos et al. 2011b) and full-scale (Lignos et al. 2013) steel moment frames during
shake table tests.

3.4.6 Definition of collapse


Only the sidesway mode of collapse is considered; other collapse modes such as vertical
collapse due to loss of axial capacity in a column or in beam-column connections are
neither simulated nor considered. In this study collapse is defined as dynamic instability
where the lateral displacement of the structure increases without bounds. The equation of
motion is solved using Newmarks average acceleration method (Newmark 1959) with a
Krylov-Newton solution algorithm (Scott and Fenves 2010). Numerical convergence
issues are handled via a variable time step function by Lignos that reduces the time step
when convergence issues are detected, and if problems persist the solution algorithm is
changed (using the current tangent instead of the initial tangent, or vice-versa) and/or the
time step is further reduced. As a last resort, ten steps with a time step of 10-4 s are solved
using the Newton-Raphson algorithm (Scott and Fenves 2010) with a fixed number of
iterations (twenty) without checking for convergence. Handling numerical convergence
Chapter 3: Collapse Risk Assessment Case Study 49

issues in this manner permits the simulation of very large inter-story 1 drift ratios (IDRs).
As a practical limit, the analysis is stopped when the IDR in any story exceeds 0.2 rad.
This number is based on engineering judgment as a structure is not expected to reach this
level of IDR without collapsing even if it can be simulated numerically. Examination of
response history results showed that the IDR was typically increasing without bounds
when it reached 0.2 rad. Examination of response history results also revealed that the
collapse intensity is not very sensitive to this maximum IDR limit imposed, provided it is
at least 0.1 rad. An example of the relative insensitivity of the collapse intensity to the
maximum IDR limit is illustrated in Figure 3.6, which shows the first-story (the
controlling story) drift ratio (IDR1) history using a record from the 1999 Chi-Chi, Taiwan
earthquake for different ground motion intensities. For an IDR limit of 0.1 the collapse
intensity is between 0.575 g and 0.588 g while for an IDR limit of 0.2 the collapse
intensity is between 0.588 g and 0.594 g, meaning that the change in collapse intensity is
less than 0.02 g for this ground motion.

Figure 3.6: First-story drift ratio (IDR1) history under a Chi-Chi ground motion for
various ground motion intensities, illustrating how the IDR limit affects the collapse
intensity.

1
As noted by the late Professor Helmut Krawinkler, the term should be intra-story and not inter-story
because the drift is measured between points at the top and bottom of a story, not between stories.
Nevertheless the term inter-story is used here because it is commonly used in the structural engineering
literature.
Chapter 3: Collapse Risk Assessment Case Study 50

3.5 Simulation and collapse risk results


This section presents simulation results for the case study including a pushover curve,
results from response history analysis including a collapse fragility curve, calculation of
the mean annual frequency of collapse (c), and a c deaggregation curve.

A pushover of the structure is conducted with lateral loads applied at the floor levels
consistent with the equivalent static lateral load pattern specified by ASCE 7-10 (ASCE
2010, Section 12.8.3). The structure forms a mechanism involving the lower three stories
under this type of loading as illustrated schematically in Figure 3.7(a). The pushover
curve is shown in Figure 3.7(b), which shows the base shear normalized by the weight of
the structure as a function of the roof drift ratio (roof displacement normalized by the
roof height) and as a function of IDR1. The drift ratios of the roof and first story differ
after approximately 0.02 radians because the first story is involved in the mechanism
whereas not all stories along the height of the building are involved in the mechanism
(i.e., the top story is not involved in the mechanism). The kink in the pushover curve at
drift ratios of approximately 0.08 radians for the roof and 0.10 radians for the first story is
due to elements reaching their residual strengths.

(a) (b)

Figure 3.7: Pushover analysis of case study structure: (a) schematic of collapse
mechanism; (b) pushover curve.

The performance of the structure is evaluated for each ground motion using incremental
dynamic analysis (IDA) (Vamvatsikos and Cornell 2002) in which nonlinear response
Chapter 3: Collapse Risk Assessment Case Study 51

history analyses (RHAs) are repeated using increasing levels of ground motion intensity
until dynamic instability is produced. The minimum intensity that results in dynamic
instability is called the collapse intensity. For this case study, the intensity of the ground
motion is measured by the 5%-damped pseudo-spectra acceleration ordinate at the
fundamental period of vibration of the structure, Sa(T1), and for each record the collapse
intensity is found to the nearest 0.01 g. Because only two-dimensional analyses are
performed, the two horizontal components of a ground motion are analyzed separately.
This results in two collapse intensities for each of the 137 ground motion pairs in the set
for a total of 274 collapse intensities. Because having a large sample of ground motions is
a priority for other chapters, each component is considered an independent event and the
collapse fragility of the structure is computed with all of the 274 collapse intensities (i.e.,
a controlling component is not used). IDA curves for each of the 274 records are shown
in Figure 3.8, where each curve denotes the intensity of the ground motion measured as
Sa(T1) as a function of the maximum IDR experienced in any story. The intensity at
which a curve becomes horizontal indicates the collapse intensity as an infinitely small
increase in intensity produces an infinitely large increase in displacement. The thick
black line in Figure 3.8 denotes the median Sa(T1) conditioned on the value of maximum
IDR. The median line confirms that, for this case study, the collapse intensity is generally
not sensitive to the IDR limit imposed provided it is greater than 0.1 rad, as the difference
in the median Sa(T1) between IDRs of 0.1 rad and 0.2 rad is approximately 0.02 g.

Figure 3.8: IDA curves for each of the 274 records used in the case study. The thick black
line denotes the median Sa(T1) conditioned on the value of maximum IDR.
Chapter 3: Collapse Risk Assessment Case Study 52

The collapse fragility curve, which shows how the probability of collapse increases with
ground motion intensity, is shown in Figure 3.9 along with the collapse intensities from
all 274 records. The ordinate associated with the collapse intensity of a given record in
Figure 3.9 represents the fraction of records whose collapse intensities are less than or
equal to that of the given record. The collapse fragility curve is obtained by fitting a
lognormal distribution to the 274 collapse intensities using the method of moments. Note
that although other plotting positions (e.g., Hazen, Weibull, Blom, etc.) and other fitting
methods (e.g., linear regression on the data plotted on lognormal probability paper,
maximum likelihood, etc.) could have been used they lead to practically identical results
for large samples such as the one used here. This results in a lognormal curve defined by
a geometric mean or median value (these two quantities are equal for the lognormal
distribution) of Sa(T1) = 1.03 g and a lognormal standard deviation of ln = 0.39. The
lognormal standard deviation is commonly called the dispersion and denoted as (e.g.,
ATC 2009a). The dispersion of 0.39 is due exclusively to record-to-record variability as
it was computed for a single analytical model of the building (i.e., it does not include
modeling uncertainty).

Figure 3.9: Collapse fragility for the case study including individual collapse intensities
and various probability distributions fit to the data. The lognormal distribution is selected
as the best fit to the data. The lognormal curve is defined by a median collapse intensity
of Sa(T1) = 1.03 g and a lognormal standard deviation of ln = 0.39.
Chapter 3: Collapse Risk Assessment Case Study 53

The lognormal distribution is typically used to characterize the distribution of collapse


intensities, and previous studies (e.g., Ibarra and Krawinkler 2005; Bradley and Dhakal
2008; Ghafory-Ashtiany et al. 2011) that conducted goodness-of-fit tests have concluded
that the collapse fragility curve can be assumed to follow a lognormal distribution. For
reference, collapse fragility curves using the Weibull and gamma probability distributions
are also shown in Figure 3.9. These distributions are also fit to the collapse intensities
using the method of maximum likelihood. The Weibull fit fails the Kolmogorov-Smirnov
goodness-of-fit test (Benjamin and Cornell 1970) at the 5% confidence level; however,
both the lognormal and gamma fits pass. The lognormal distribution is visually judged to
be a better fit to the data than the gamma distribution, especially at the intensities
contributing most to the collapse risk, and therefore the former is used in all subsequent
representations of the collapse fragility curve.

The mean annual frequency of collapse c is computed by integrating the product of the
collapse fragility curve and the slope of the seismic hazard curve over all ground motion
intensities as shown in Equation 2.3. This results in c = 3.9x10-4, which has a similar
order of magnitude to values obtained in a study by Haselton and Deierlein (2008) in
which they assessed the seismic performance of modern reinforced concrete moment
frame buildings located at the same site in southern California and found that c values
were on the order of 10-5 to 10-4 for these buildings. The c for the case study building
corresponds to a 1.9% probability of collapse of over 50 years (Pc,50) using Equation 2.5.
It should be noted that this value of Pc,50 is approximately twice the target value for
structures designed according to ASCE 7-10 (ASCE 2010); however, the collapse risk of
the case study structure was computed without directly considering how the spectral
shape of the records used may bias the computed risk with respect to what would be
computed using records whose spectral shapes are consistent with what is expected at the
site based on a PSHA deaggregation. The effects of spectral shape on structural response
estimates are examined in detail in Chapter 5, and the effects of spectral shape on the
computed collapse risk of the case study structure are examined in Section 5.7 and
Section 6.4. Effects of ground motion duration on the structural response estimates are
not considered here, although recent research suggests that duration can influence the
Chapter 3: Collapse Risk Assessment Case Study 54

collapse risk estimate (Chandramohan et al. 2013). This is a particularly important


consideration for sites near zones capable of generating large magnitude events where
long duration motions are expected to occur.

Figure 3.10 shows the components involved in computing c including the collapse
fragility curve (Figure 3.10(a)), the slope of the seismic hazard curve (Figure 3.10(b)).
The product of these two components is shown in Figure 3.10(c), which is called the c
deaggregation curve as it shows each intensitys contribution to c. As shown in Figure
3.10, intensities corresponding to the lower half of the collapse fragility curve dominate
the structures collapse risk. The most significant contribution to c is from Sa(T1) = 0.74
g (Figure 3.10(c)), which corresponds to a 20% probability of collapse (Figure 3.10(a)).
For the case study, 65% of c comes from intensities less than the median collapse
intensity (Sa(T1) = 1.03 g). The dominance of intensities less than the median collapse
intensity is driven by the steep slope of the seismic hazard curve at these intensities,
which outweighs the corresponding small probabilities of collapse. In other words, even
though large levels of intensity (e.g., > 1.5 g) lead to very large probabilities of collapse
(larger than 80% for this structure), these large levels of ground motion intensity are very
rare and have much smaller frequencies of exceedance than those in the lower portion of
the collapse fragility curve, and therefore they do not end up contributing significantly to
c. Meanwhile, ground motion intensities in the lower half of the collapse fragility curve,
even though they have much smaller probabilities of collapse, occur significantly more
frequently (by one or more orders of magnitude) than those associated with large
probabilities of collapse and therefore lead to the largest contributions to c.

These findings related to which intensities contribute most significantly to c are


consistent with other studies that examined c deaggregations. A summary of these
studies was presented by Eads et al. (2012) and is repeated here. Ibarra and Krawinkler
(2005) examined five single-degree-of-freedom (SDOF) systems designed for the Los
Angeles area with periods ranging from 0.1 s to 2.0 s and found that the maximum
contributions to c came from intensities associated with probabilities of collapse around
20%. Haselton and Deierlein (2008) showed that intensities around 60% of the median
Chapter 3: Collapse Risk Assessment Case Study 55

collapse intensity provided the maximum contributions to c for two 4-story reinforced
concrete frames also designed for the Los Angeles area. Deaggregations of c for
structures located throughout New Zealand presented by Bradley and Dhakal (2008) also
showed that the most significant contributions to c came from intensities smaller than the
median collapse intensity.

Figure 3.10: Case study collapse risk assessment: (a) collapse fragility curve; (b) slope of
the seismic hazard curve; (c) c deaggregation curve.

3.6 Influence of the seismic hazard curve


This section examines the influence of the seismic hazard curve on the collapse risk,
specifically how the shape of the hazard curve affects the shape of the c deaggregation
and the relative value of c. To accomplish this, seismic hazard curves are obtained for
six locations throughout the United States (US). These locations, which are defined in
Table 3.2, comprise three sites in the western US and three sites in the central and eastern
US. The c deaggregations are computed for each location using the response history
Chapter 3: Collapse Risk Assessment Case Study 56

results from the case study in Section 3.5. The material in this section appeared in a paper
by Eads et al. (2012), although notable differences include use of: (1) a different hazard
curve due to the availability of updated hazard data since the paper was submitted; and
(2) a different ground motion set, which produces small changes in the collapse fragility
curve.

Table 3.2: Locations used in case study.


Western US Central & Eastern US
Site Zip Code Site Zip Code
Los Angeles, CA 90015 Charleston, SC 29401
San Francisco, CA 94111 Memphis, TN 38103
Seattle, WA 98104 St. Louis, MO 63101

The seismic hazard curves were again obtained from the USGS Hazard Curve
Application (USGS 2012a). Because data was unavailable for NEHRP Site Class D for
the central and eastern locations, the seismic hazard data for the border of NEHRP Site
Classes B and C is used for all locations in the interest of consistency. This is deemed
reasonable as the primary objective here is to examine how differences in the seismic
hazard curves affect the c deaggregation and the relative value of c. The seismic hazard
curves are interpolated to T1 using the method described in Section 3.2. The resulting
seismic hazard curves are shown in Figure 3.11, where the lines marked 10/50 and
2/50 denote the hazards with 10% and 2% probabilities of exceedance in 50 years,
respectively. Events associated with these hazard levels are often referred to as the
Design Basis Event (DBE) and Maximum Considered Event (MCE), respectively.
Chapter 3: Collapse Risk Assessment Case Study 57

Figure 3.11: Seismic hazard curves for Sa(T1) for various locations in the US.

The c values for the various locations are shown in Table 3.3. These values are not the
focus of this section and are tabulated here for illustrative purposes only as the structure
and ground motions used for computing the collapse fragility function are not necessarily
representative of what is expected at each location. The focus, rather, is only on how
differences in the seismic hazard curves affect the relative values of c. As the same
collapse fragility curve is used for each site, differences in c are exclusively due to
differences in the seismic hazard curves. Because the seismic hazard curves of Los
Angeles and St. Louis are very different (Figure 3.11), one would expect c to be very
different for these sites. This is the case as c for Los Angeles is over forty times larger
than c for St. Louis (Table 3.3). Using the same logic, one might also expect c for
Seattle and Memphis to be very different; however, these two locations have nearly
identical values of c (Table 3.3) even though the seismic hazard in Seattle is greater at
both the 10/50 and 2/50 levels (Figure 3.11). Intensities associated with the 2/50 level are
similar for Memphis and Charleston, with the intensity for Memphis being slightly larger
(Figure 3.11). While c is on the same order of magnitude for these sites, c for Memphis
is actually 34% smaller than c for Charleston (Table 3.3). This is due to differences in
the seismic hazards at intensities other than the 2/50 level. The previous results
emphasize the importance of considering all intensities that contribute to the collapse risk
Chapter 3: Collapse Risk Assessment Case Study 58

as opposed to some of the collapse metrics described in Section 2.4 that are based on only
a single level of ground motion intensity. Recognizing that the collapse risk of a structure
designed based on the spectral ordinate at the 2/50 level varies with its location due to
differences in the shapes of seismic hazard curves (Luco et al. 2007), ASCE 7-10 (ASCE
2010, Chapter C11) updated the probabilistic portions of its seismic design maps with
risk-targeted spectral ordinates aimed at having more uniform collapse risks for structures
in different locations.

Table 3.3: c values for case study.


Western US Central & Eastern US
Site c [10 ]
5 Site c [105]
Los Angeles, CA 5.8 Memphis, TN 2.5
San Francisco, CA 6.0 Charleston, SC 3.8
Seattle, WA 2.4 St. Louis, MO 0.1

The c deaggregations for all sites are presented in Figure 3.12(e,f) along with the
collapse fragility curve (Figure 3.12(a,b)) and the slopes of the seismic hazard curves
(Figure 3.12(c,d)). Plots on the left pertain to the western US sites while plots on the right
pertain to the central and eastern US sites. The deaggregations show that for all sites,
despite significant differences in the seismic hazard curves, the largest collapse risk
contributions (the peak of each deaggregation curve in Figure 3.12(e,f)) occur at
intensities less than the median collapse intensity of Sa(T1) = 1.03 g. In fact, the intensity
with the maximum contribution to c is in the lower third of the collapse fragility curve
for all sites. This is the result of the slopes of the seismic hazard curve being typically
much larger (sometimes by orders of magnitude) for intensities in the lower half of the
collapse fragility curve than for those in the upper half of the collapse fragility curve.
Chapter 3: Collapse Risk Assessment Case Study 59

Figure 3.12: Case study collapse risk assessment for locations throughout the US:
collapse fragility curve in (a) and (b); slopes of the seismic hazard curves in (c) and (d);
c deaggregation curves in (e) and (f). Plots on the left (i.e., (a) (c) and (e)) pertain to the
western US locations while plots on the right (i.e., (b) (d) and (f)) pertain to the central
and eastern US locations. Note the different vertical scales of the c deaggregation curves.

As previously stated the structure and ground motions used for this study are not
necessarily representative of what is expected at each location, and in general it is not
expected that the collapse fragility curve would be the same at all six locations. To test
the sensitivity of previous observations regarding which intensities have the most
Chapter 3: Collapse Risk Assessment Case Study 60

significant contributions to the collapse risk, the c deaggregations are recomputed


assuming that the collapse risk is the same at all six locations. The collapse risk is
selected to be c = 2.0x10-4, which corresponds to a 1% probability of collapse in 50
years using Equation 2.5. This value is chosen because it is consistent with the target
collapse risk of structures designed according to ASCE 7-10 (ASCE 2010, Chapter 22).
The median collapse intensity that produces a collapse risk of c = 2.0x10-4 for each site
is found through iteration assuming that the dispersion of the collapse fragility curve is
ln = 0.4, which Zareian and Krawinkler (2009) found was a typical value for generic
moment-resisting frames due to record-to-record variability (i.e., not accounting for any
modeling uncertainty). Note that determining the collapse fragility curve in this manner
means that the structure and ground motion set used previously are no longer relevant.
The assumption that the same structure and ground motions are used at all sites is not
required. The collapse fragility curve for each site simply represents a hypothetical
structure and ground motion set that produce the specified collapse fragility curve.

Collapse risk deaggregations are performed for each site, and the results are summarized
in Figure 3.13(a) for the western US sites and Figure 3.13(b) for the central and eastern
US sites. Figure 3.13 shows the cumulative contribution to c as a function of intensity,
where the cumulative contribution to c is the area under the c deaggregation between
zero and the given intensity expressed as a fraction of c (i.e., the area under the c
deaggregation between zero and an infinitely large intensity). Circles in Figure 3.13
indicate the median collapse intensity while stars indicate the intensity with the most
significant contribution to the collapse risk (i.e., the intensity at the peak of the c
deaggregation). The results presented in Figure 3.13 show that the intensity with the
greatest contribution to the collapse risk is less than the median collapse intensity for all
six sites, which is consistent with the previous example. For the western US sites shown
in Figure 3.13(a), the majority of the collapse risk comes from intensities in the lower
half of the collapse fragility curve as intensities less than the median collapse intensity
are responsible for between 58% (in the case of Seattle) and 78% (in the case of San
Francisco) of c. The converse is true for the central and eastern US sites shown in Figure
3.13(b) as intensities less than the median collapse intensity are only responsible for
Chapter 3: Collapse Risk Assessment Case Study 61

between 33% (in the case of Charleston) and 48% (in the case of St. Louis) of c. The
cumulative contribution to c at the intensity associated with the peak of the c
deaggregation was relatively consistent for all sites, however, as the cumulative
contribution at this intensity is approximately 35% for the west coast sites (Figure
3.13(a)) and approximately 30% for the central and east coast sites (Figure 3.13(b)).

Figure 3.13: Cumulative contribution to c as a function of intensity for: (a) western US


sites; (b) central and eastern US sites. All sites have a collapse risk of c = 2.0x10-4 and
dispersion on the collapse fragility curve of ln = 0.4.

Though the collapse fragility curves are not shown, it is worth noting that the intensity
associated with the peak of the c deaggregation is located in the lower third of the
collapse fragility curve (i.e., the probability of collapse at this intensity is less than 33%)
for all western US sites and for St. Louis, which in consistent with the previous example
that used the same collapse fragility curve for all sites. This does not hold for Memphis
and Charleston, however, as the intensity at the peak of the c deaggregation is associated
with collapse probabilities of 36% and 43%, respectively, for these sites.

The previous discussions about the c deaggregation curve, the intensities that contribute
most to c, and the influence of the seismic hazard curve demonstrate that the most
significant contribution to the collapse risk comes from an intensity less than the median
Chapter 3: Collapse Risk Assessment Case Study 62

collapse intensity. The cases examined here suggest that the collapse risk of the western
US sites is dominated by intensities in the lower half of the collapse fragility curve, but
this may not necessarily be true for all sites in the central and eastern US. Note that these
findings are based on case studies where the probability of collapse over a 50-year period
is less than 2%. Structures with very high collapse risks can have significant probabilities
of collapsing at ground motion intensities that occur relatively frequently, meaning that
the intensity with the greatest contribution to the collapse risk could be greater than the
median collapse intensity. Thus regardless of how well a structure is designed, it is
recommended that to obtain a good collapse risk estimate emphasis should be placed on
intensities with significant contributions to the collapse risk rather than the median
collapse intensity as is currently done (e.g., ATC 2009a). This serves as motivation for
the efficient method of estimating the collapse risk that is proposed in Chapter 4.
Chapter 4

Statistical Uncertainty in the Collapse Fragility and Mean


Annual Frequency of Collapse due to the Number of Ground
Motions Considered

This chapter is based on the following publication:


Eads, L., Miranda, E., Krawinkler, H., and Lignos, D. G. (2013). "An efficient
method for estimating the collapse risk of structures in seismic regions."
Earthquake Engineering & Structural Dynamics, 42(1), 25-41.

4.1 Overview
Previous studies have focused on uncertainty in collapse risk due to record-to-record
(RTR) variability and modeling uncertainty (e.g., Ibarra and Krawinkler (2005, Chapter
6; 2011); Zareian and Krawinkler (2007b; a); Bradley and Dhakal 2008; Liel et al. 2009).
This chapter deals with uncertainty in the collapse fragility (i.e., the probability of
collapse conditioned on ground motion intensity) and mean annual frequency of collapse
(c) associated with employing a small number of ground motions in response history
analysis (RHA). This type of uncertainty is an epistemic uncertainty because it is
reducible, in this case by using a larger sample of ground motions, and is sometimes
referred to as statistical uncertainty (e.g., Benjamin and Cornell 1970). Only
uncertainty in the collapse risk due to obtaining statistical parameters of the collapse
fragility by using a small number of ground motion records and the effect on c are
discussed here. Theoretical confidence intervals on the collapse fragility are discussed,
and uncertainty in the collapse fragility and c are quantified as a function of the number
of ground motions using the case study results from Chapter 3. Finally, a method for
estimating c based on estimating the collapse fragility curve at only two intensities is
proposed, and this method is demonstrated using the case study.

63
Chapter 4: Statistical Uncertainty in the Collapse Fragility 64

4.2 Background and literature review


According to classical statistical theory (see, e.g., Benjamin and Cornell 1970), the
standard error (SE) of a sample mean ( x ) is computed as

X
SE ( x ) = (4.1)
N

where X is the standard deviation of the random variable X of the population from
which the samples are drawn and N is the number of samples. In the context of this
chapter, N is used to denote the number of ground motions used in RHA. In practice
however, X is often unknown and is replaced in Equation 4.1 by the sample standard
deviation s X to give the estimated SE of x . As N gets large, the central limit theorem
says that the distribution of x becomes approximately normal. This facilitates the use of
confidence intervals on x computed using the normal probability distribution. However,
if the parameter being estimated is not x , the SE of the estimated parameter cannot be
computed using Equation 4.1 and in most cases a formula for estimating the SE of a
parameter other than x does not exist (Efron and Tibshirani 1993). Furthermore, the
central limit theorem, which facilitates approximating the estimated parameter as
normally distributed, technically only applies when: (1) N is large; and (2) the estimated
parameter is the sum of independent random variables. Estimates of the mean fall under
the central limit theorem because the sample mean is simply the sum of random variables
divided by a constant (i.e., the number of samples, N).

The bootstrap procedure developed by Efron (1979) provides a way to estimate the SE of
almost any statistic or estimator and can also be used to construct confidence intervals on
the estimated parameter without the limitations of the central limit theorem. A detailed
description of the bootstrap procedure is provided by Efron and Tibshirani (1993), but the
essentials are as follows: given a set of N observations, an arbitrarily large number of
bootstrap samples are generated, where each bootstrap sample contains N observations
randomly sampled, with replacement, from the original observations. The desired
parameter (e.g., x ) is computed for each bootstrap sample, and the bootstrap estimate of
Chapter 4: Statistical Uncertainty in the Collapse Fragility 65

SE is computed as the standard deviation of the bootstrap replications of the desired


parameter. The bootstrap replications can also be used to construct confidence intervals
(see Efron and Tibshirani 1993 for details).

A number of previous studies have examined how the number of ground motions used in
RHA affects the uncertainty in structural performance estimates, and many of these
studies employed the bootstrap procedure. These studies have typically focused on
uncertainty in the mean or median value 1. In particular, many studies looked at the
variability in the median capacity (i.e., the horizontal variability of one point on the
collapse fragility) as opposed to the variability of the probability of collapse conditioned
on ground motion intensity, P(C | IM), (i.e., the vertical variability of all points on the
collapse fragility). For example, Jalayer (2003, Chapter 4) estimated the statistical
uncertainty in the median estimate of the peak interstory drift ratio (IDR) conditioned on
ground motion intensity as a function of N. She found that the SE in the median estimate
was approximately equal to the value obtained using Equation 4.1. Vamvatsikos and
Cornell (2006) compared the error in estimating the 16th, 50th, and 84th percentiles of
incremental dynamic analysis (IDA) curves based on demand and capacity using a set of
10 records versus using a set of 30 records. The ratio of error using 10 records versus 30
records was, on average, 1.7, which is what is expected based on Equation 4.1 (i.e., (30)
/ (10) 1.7); however, when considering a specific type of force-displacement curve
(i.e., bilinear, trilinear, or quadrilinear) this error ratio ranged from 1.4 to 2.5. Aslani and
Miranda (2004) studied the number of ground motions required to capture and median,
logarithmic standard deviation, and mean annual frequency of exceedance of peak IDRs
and peak floor accelerations. They found that the number of ground motions required to
estimate both the median IDR for a given ground motion intensity level and the mean
annual frequency of exceeding a given IDR level within a certain error range (e.g., 10%
error) increased with the intensity level and the IDR level, respectively. They found that
1
The lognormal distribution is commonly used in structural engineering to describe the distribution of
structural responses (e.g., interstory drift ratios, floor accelerations, collapse intensities, etc.). For a
lognormally distributed random variable Y (i.e., the natural logarithm of Y, denoted as ln(Y), is normally
distributed), the median of Y is equal to the geometric mean of Y, which is the exponential of the mean of
ln(Y). Because of this equality, the median of Y is often used to characterize the lognormal distribution or
report results as it has the same units as the response parameter of interest (as opposed to the mean of
ln(Y)).
Chapter 4: Statistical Uncertainty in the Collapse Fragility 66

the number of ground motions required to estimate the lognormal dispersion of the
median IDR for a given ground motion intensity level within a certain error range was
nearly independent of intensity level, however.

In a study related to the procedure proposed by Eads et al. (2013), which is discussed
later in this chapter, Baker (2013) estimated the collapse fragility curve by estimating
P(C | IM) at three intensity levels. In a case study involving Monte Carlo simulation of
1,000 realizations of the collapse fragility curve, he found that the standard deviation of
the median collapse intensity reduced from 0.078 when using 20 ground motions per
intensity to 0.056 when using 40 ground motions per intensity. This reduction is also
consistent with what is expected based on Equation 4.1 (i.e., 0.078/0.056 (40) / (20)
1.4); however, it is not clear that the assumptions behind Equation 4.1 are satisfied for
this case because the median collapse intensity is estimated by on fitting a lognormal
distribution to the three P(C | IM) data points and taking the median of that distribution as
opposed to computing the median from a set of 20 or 40 collapse intensities.
Furthermore, the standard deviation is computed from 1,000 samples of the median
collapse intensity in both cases. The number of ground motions used is directly involved
in only the estimation of P(C | IM), not the subsequent fitting of the collapse fragility
curve.

Baker (2013) also studied how the number of ground motions affected the estimate of c.
The collapse fragility curves were estimated as previously described and were combined
with a seismic hazard curve to yield 1,000 estimates of c for each case considered. When
using a seismic hazard curve for Los Angeles, Baker found that the coefficient of
variation (COV), which is the standard deviation normalized by the mean, of c reduced
from 0.27 to 0.20 when the number of ground motions used per intensity increased from
30 to 45. When looking at two other seismic hazard curves, he found the COV decreased
from 0.20 to 0.15 in one case and from 0.55 to 0.33 in the other case. Assuming that the
mean values of c for the 30 and 45 ground motion cases are approximately equal for a
given seismic hazard curve, comparing the COVs of c is approximately equivalent to
comparing the standard deviations of c. Equation 4.1 predicts that the ratio of standard
Chapter 4: Statistical Uncertainty in the Collapse Fragility 67

error would be (45) / (30) 1.2, whereas the COV ratios for each hazard curve case
are approximately 1.4, 1.3, and 1.7. It is not clear that Equation 4.1 applies to this case for
the reasons discussed in the preceding paragraph; however, the examples presented here
demonstrate that the uncertainty in collapse risk metrics does not always decrease
proportionally to 1/N.

Shome et al. (1998) state that the number of records required to estimate the median
structural response Y within a factor of X (i.e., within a range of Y(1 X) ) with 95%
confidence can be approximated as

4 ln2 Y
N (4.2)
X2

where ln Y is the logarithmic standard deviation of Y, which is commonly called the


dispersion and denoted as . As shown by Huang (2008; Appendix C), the former
approximation was derived from the formula for determining the number of records
required to estimate the median of a lognormally distributed random variable Y within a
factor of X with (1-)% confidence

2
1
1 ln Y
2
N = (4.3)
ln ( X + 1)

where -1( ) is the inverse of the standard normal cumulative distribution function and
ln( ) is the natural logarithm.

A similar formula based on the central limit theorem is often used to estimate the number
of Monte Carlo simulations required to compute the probability of failure (pf) for a given
COV of the pf estimator (see, e.g., Rubinstein and Kroese 2008)
Chapter 4: Statistical Uncertainty in the Collapse Fragility 68

1 p f
N (4.4)
COV 2 p f

To obtain a reasonable estimate of the median structural response when performing


intensity-based assessments, Section 5.2.2 of FEMA P-58 (ATC 2012) recommends
using at least 7 pairs of ground motions and notes that 11 pairs can provide a reasonable
estimate when the spectral shape of the motions does not match the target spectrum. The
recommendation of 11 ground motion pairs is derived from Equation 4.3 and is aimed at
predicting the median response within 20% of the true median with 75% confidence
assuming that ln Y = 0.52, which was the largest dispersion on structural responses
obtained in case studies (Huang 2008; Appendix C). When predicting the median
collapse intensity via IDA or estimating the collapse fragility, Section 6.2.3 of FEMA P-
58 recommends using at least 20 ground motion pairs, although Section J.4 notes that
limited studies showed that using 11 ground motion pairs (each rotated 90 degrees to
produce 22 ground motion sets) was sufficient to predict the median collapse intensity.
To accurately predict the dispersion in the median results due to RTR variability, Section
5.2.5 of FEMA P-58 states that at least 30 ground motion pairs are generally needed.
When assessing the error in estimating the dispersion of median IDR conditioned on
ground motion intensity, Aslani and Miranda (2004) found that using 60 records (i.e., 30
ground motion pairs) led to dispersion estimates within 20% of the exact value with
95 confidence for a case study structure. They also observed that the confidence intervals
on the estimate can be non-symmetric, particularly when using a small number of ground
motions. In lieu of conducting many RHAs to estimate this dispersion, Section J.5 of
FEMA P-58 recommends using = 0.45 or using another pre-defined value based on the
structural period and geographic region to estimate the dispersion due to RTR variability.

FEMA P-58 usually recommends using a larger number of ground motions in RHA for
estimating the mean/median structural response compared to other guidelines. For
example, Chapter 16 of ASCE 7-10 (ASCE 2010) requires a minimum of seven ground
motions be used in order to use the average of the peak member forces and story drifts for
design. Using a minimum of three ground motions in RHA is permitted, except that the
Chapter 4: Statistical Uncertainty in the Collapse Fragility 69

maximum (not the average) of the peak response values must be used for design. It
should be noted, however, that FEMA P-58 is intended for performance assessment and
evaluation of losses, and this implies a performance-based/higher level of design than
many guidelines, which are primarily intended to provide life safety in a design event.

4.3 Confidence intervals on the probability of collapse conditioned on


intensity
This section discusses theoretical confidence intervals on the probability of collapse
conditioned on ground motion intensity, P(C | IM), which represents an ordinate on the
collapse fragility curve. This ordinate is often estimated as the fraction of ground motions
that produce collapse in the structure when the ground motions are scaled to a certain
intensity level. Since collapse and non-collapse events are mutually exclusive, for a given
ground motion intensity this is equivalent to the structural response adopting a two-
valued (binary) random variable (e.g., collapse = 1 and no collapse = 0). Therefore,
estimating ordinates of the collapse fragility curve is equivalent to estimating the
proportion of ground motions that produces collapse from the total number of ground
motion in the sample. The latter can be estimated using the binomial probability
distribution which is a discrete probability distribution whose cumulative distribution
function (CDF) is given by

x
P( X x) =

k =0
N!
( N k )!k!
p k (1 p ) N k (4.5)

where X is the binomial random variable, x is number of successes (collapses), N is the


number of trials (ground motions), and p is the probability of success in a single trial.

The level of uncertainty in estimating the ordinate of the collapse fragility at a given
ground motion intensity level im, P(C | IM=im), is a function of the number of ground
motions used to make the estimation and the observed binomial proportion at intensity
im which corresponds to the fraction of ground motions producing collapse. The latter is
computed as
Chapter 4: Statistical Uncertainty in the Collapse Fragility 70

n
P (C | IM ) p im = im (4.6)
N

where nim is the number of ground motions producing collapse at ground motion intensity
level im and N is the total number of ground motions used in the estimation. One of the
oldest and most commonly used methods to estimate the confidence intervals of a
binomial proportion are the Wald confidence intervals (see, e.g., Agresti and Coull 1998)
given by

p im (1 p im )
[PC ,LB , PC ,UB ] = p im z / 2 N
(4.7)

where PC,UB and PC,LB are the 100(1-)% upper and lower bound confidence intervals of
the probability of collapse at ground motion intensity level im, and z/2 is the 1-/2
quantile of the standard normal distribution. Examination of Equation 4.7 indicates that
these confidence intervals are symmetrically spaced around pim and that the uncertainty
is reduced as the sample size (the number of ground motion records) is increased. Wald
confidence intervals are based on the asymptotic normality of the sample proportion and
therefore provide a good estimate of the confidence interval for large sample sizes when
the sample proportion is not very close to 0 or 1.

An exact method for obtaining binomial confidence intervals was developed by Clopper
and Pearson (1934); however, Agresti and Coull (1998) showed that even though the
Clopper-Pearson confidence intervals are exact, they tend to be conservative because the
actual coverage probability can be much larger than the nominal confidence level unless
N is quite large. Agresti and Coull judged these intervals to be inappropriate for statistical
practice and recommended using the Wilson score confidence intervals (Wilson 1927),
which are appropriate for all sample sizes and sample proportions and are given by
Chapter 4: Statistical Uncertainty in the Collapse Fragility 71


z2 pim (1 pim ) z2 / 2
pim + / 2 z2 / 2 +
2N N 4N 2
[PC ,LB , PC ,UB ] = (4.8)
z2 / 2
1+
N

Figure 4.1 shows the 95% Wilson confidence intervals computed using Equation 4.8 for
the probability of collapse as a function of the fraction of ground motions producing
collapse ( pim ) and the number of ground motions N. The confidence intervals are shown
for N = 10, N = 40 (a number commonly used to compute the collapse fragility curve,
e.g., Ibarra and Krawinkler 2005; Zareian and Krawinkler 2009; Lignos and Krawinkler
2012a), and N = 274 (the number of ground motions used in the case study from Chapter
3). It can be seen in Figure 4.1 that when using a small number of ground motions the
uncertainty in the ordinates of the collapse fragility is very high. For example, if one uses
a set of only 10 ground motions and when scaled to a certain ground motion intensity one
computes a probability of collapse of 40% (i.e., pim = 0.40), there is 95% confidence that
the actual probability of collapse at that intensity level (i.e., the value computed with an
infinitely large set of ground motions) could be at low as 17% or as high as 69%. As
shown in Figure 4.1, the uncertainty in the probability of collapse decreases as the
number of ground motions increases, meaning that the uncertainty in the collapse
fragility is reducible and therefore, by definition, is an epistemic uncertainty.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 72

nim
p im =
N
Figure 4.1: 95% binomial confidence intervals for P(C | IM): Wilson score confidence
intervals as a function of pim for different N.

Figure 4.1 also shows that the confidence intervals are not symmetric. For small values of
pim , which are of particular interest as they typically have larger contribution to c as

discussed in Chapter 3, there is a chance that the true probability of collapse is


significantly greater than the estimated value. This case is of particular engineering
interest as RHAs are often performed at intensities where the probability of collapse is
expected to be small and because intensities with small probabilities of collapse can have
significant contributions to the collapse risk as discussed in Sections 3.5 and 3.6. As an
example, assume that RHA results obtained using a sample of 10 ground motions show a
10% probability of collapse at a given intensity level (i.e., pim = 0.10). The upper bound
of the 95% confidence interval for the true probability of collapse at that intensity is 40%
(30% greater, in absolute terms, than the estimated probability) while the lower bound is
only 2% (8% less, in absolute terms, than the estimated probability). As shown in Figure
4.1, both the width and the asymmetry of the confidence interval at a given value of pim
reduce as the number of ground motions increases.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 73

4.4 Case study of statistical uncertainty in the collapse risk


This section uses the results from the case study presented in Chapter 3 to illustrate the
statistical uncertainty in the collapse risk. To demonstrate the variability in pim estimates
as a function of the number of ground motions used in the estimation, N ground motions
were randomly selected with replacement from the 274 motions used in the case study
and were used to compute pim as shown in Equation 4.6. This process was repeated 100
times for each N, and the 100 realizations of pim for Sa(T1) = im = 0.74 g are shown in
Figure 4.2. From the case study of Los Angeles discussed in Section 3.5, the intensity of
0.74 g had the maximum contribution to c and yielded P(C | im = 0.74 g) = 20%, which
is denoted by the dashed horizontal line in Figure 4.2. Because this intensity has the
largest contribution to c, a good estimation of c depends significantly upon the quality
of the pim estimate in the neighborhood of this intensity. Figure 4.2(a) shows that
significant variability in pim exists for small N. In sixteen of the 100 realizations for N =
10, the ground motion sample produced pim (a probability of collapse) of at least 40%, or
at least two times greater than the probability of 20% computed with 274 motions. In
other words using a small number of ground motions such as N = 10 could lead to
significant chances of producing large overestimations of the probability of collapse by
more than two times the one you would compute using many ground motions. Similarly,
when using a single set of N = 10 you could be computing a null probability of collapse
(none of the ten records produce collapse) whereas with a large ground motion set one
would compute values close to 20%. The former situation occurred in eight of the 100
trials, indicating there was also a significant probability of largely underestimating the
probability of collapse when N = 10. As N increases the variability of the pim realizations
decreases, which indicates increased confidence in the estimate.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 74

Figure 4.2: Realizations of pim for im = 0.74 g with different N: (a) N = 10; (b) N = 40;
(c) N = 274.

4.4.1 Confidence intervals on the collapse fragility curve


To quantify the uncertainty in the collapse fragility curve as a function of N,
bootstrapping techniques were used to construct 95% confidence intervals on pim using
the percentile method (Efron and Tibshirani 1993). For a given N and im, 1,000
realizations of pim were obtained using the procedure described in the preceding
paragraph. The number 1,000 was used to obtain relatively smooth and stable confidence
intervals. The confidence intervals for the case studys collapse fragility curve are shown
in Figure 4.3 for different N along with simulation results from the case study (i.e., the
empirical CDF of collapse intensities). It should be noted that pim was estimated at
discrete im levels and that the confidence interval for a given N was constructed by
connecting the appropriate percentile (i.e., 2.5% or 97.5%) estimates of pim with a
straight line. Figure 4.3 shows that significant uncertainty exists in the collapse fragility
when using a small N, which can lead to relatively large errors in estimating the collapse
fragility. For example the 95% confidence interval on the probability of collapse at an
intensity of 1.03 g (the median collapse intensity from the case study) was [0.20, 0.80]
when using N = 10, which gives a ratio of 4.0 between the upper and lower confidence
limits. This ratio decreases to 1.9 for N = 40 and 1.3 for N = 274.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 75

Figure 4.3: Collapse fragility curve with 95% confidence intervals on P(C | IM) based on
simulations using the case study results shown for different N.

4.4.2 Confidence intervals on the mean annual frequency of collapse


The uncertainty in c due to N was quantified by the same confidence interval techniques
used for the collapse fragility curve (i.e., the bootstrap percentile method). For each
sample, a collapse fragility curve was generated by fitting a lognormal distribution to the
collapse intensities of N randomly selected ground motions, and c was computed with
this collapse fragility curve. To obtain smooth confidence intervals 1,000 sample sets
were used for each N. The 95% confidence intervals for c computed for the case study
are presented in Figure 4.4, which shows that significant uncertainty in c exists when
using a small N. The ratio between the upper and lower confidence intervals for c is 1.3
when N = 274, and it increases to 2.0 when N = 40 and to 4.0 when N = 10. Figure 4.4
also shows that the mean value of c is relatively stable for N > 15 (i.e., the mean of
samples does not vary significantly as a function of N) and that the confidence intervals
are approximately symmetric about the mean value for N > 50.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 76

Figure 4.4: 95% confidence intervals for c as a function of N.

4.5 Proposed method for estimating the mean annual frequency of


collapse
The computational effort in collapse risk assessment is concentrated in performing
nonlinear RHAs to obtain the collapse fragility curve. Obtaining the collapse intensity for
each ground motion is very computationally demanding as it requires conducting RHAs
at increasing levels of intensity until collapse is produced and then conducting additional
RHAs within the last interval of intensities to find the collapse intensity within a certain
tolerance. The computational effort depends, among other things, on: (a) the number of
ground motions used in collapse risk assessment; (b) the duration of each ground motion;
(c) the intensity increment(s) used in the incremental dynamic analyses; and (d) the
tolerance on collapse intensity. However, if one assumes that the collapse fragility is well
represented by a lognormal distribution, which several studies (Ibarra and Krawinkler
2005; Bradley and Dhakal 2008; Ghafory-Ashtiany et al. 2011) conclude is a reasonable
assumption, then it is not necessary to determine the collapse intensity for each ground
motion. One can construct a relatively good estimate of the entire collapse fragility curve
by conducting analyses at only two intensity levels, that is, by obtaining two points on the
collapse fragility curve. This section proposes a method for estimating c which, by using
many ground motions at specific intensities, provides more reliable results and at the
same time reduces the computational effort involved by significantly reducing the
Chapter 4: Statistical Uncertainty in the Collapse Fragility 77

required number of RHAs. To obtain a good estimate of c, the proposed method


concentrates on obtaining good estimates of the collapse fragility curve at intensities with
significant contributions to c.

4.5.1 Method overview


The proposed method was developed based on observations presented in Sections 3.5 and
3.6 regarding deaggregation of collapse risk and those discussed in this chapter regarding
statistical uncertainty when using a small number of ground motions. Given a structure
and a seismic hazard curve, the proposed method to estimate c uses a rough initial
estimate of the collapse fragility curve to identify intensities with significant
contributions to c, updates the fragility curve based on the results of RHAs at the
significant intensities, and updates the estimate of c.

4.5.2 Description of method


The proposed method to estimate c is summarized as follows:
1. Obtain an initial estimate of the collapse fragility curve assuming a lognormal
probability distribution by estimating the median collapse intensity () and the
dispersion (i.e., the logarithmic standard deviation ) using an approximate
method such as those discussed in Section 2.5 or any other approximate method.
An accurate estimation of the collapse fragility curve is not necessary.
2. Using the estimate of the collapse fragility curve from Step 1 and the seismic
hazard curve at the site, calculate c using Equation 2.4 and obtain the c
deaggregation as presented in Section 2.4.4. Identify the intensity at which the
cumulative contribution to c is approximately 90% (i.e., the intensity at which the
area under the c deaggregation curve between zero and this intensity is 90% of
c). Call this intensity IM1.
3. Conduct RHAs using ground motions scaled to intensity level IM1 and estimate
P(C | IM1) as the fraction of motions causing the structure to collapse. The
motions should be consistent with the magnitude, distance, focal mechanisms, and
site conditions from the PSHA deaggregation at the hazard level associated with
Chapter 4: Statistical Uncertainty in the Collapse Fragility 78

IM1. One can also take into account . Further information on record selection can
be found in Baker and Cornell (2006a).
4. Using the estimate of P(C | IM1) in Step 3 and the rough estimate of the
dispersion on the collapse fragility curve from Step 1, obtain an estimate of the
collapse fragility curve assuming a lognormal probability distribution. Use this
estimate of the collapse fragility curve and the seismic hazard curve at the site to
calculate c using Equation 2.4 and obtain the c deaggregation as presented in
Section 2.4.4. Identify the intensity at which the cumulative contribution to c is
approximately 20% 2 (i.e., the intensity at which the area under the c
deaggregation curve between zero and this intensity is 20% of c). Call this
intensity IM2.
5. Repeat Step 3 using IM2 to obtain P(C | IM2). Note that the ground motions used
at IM2 are not necessarily the same as those used at IM1.
6. Assuming a lognormal distribution, update the collapse fragility curve estimate by
using the points (IM1, P(C | IM1)) and (IM2, P(C | IM2)). Note that here the
estimate of the dispersion that was used in Steps 1 and 4 is no longer used.
7. Using the collapse fragility curve from Step 6 and the seismic hazard curve,
calculate c using Equation 2.4.

If the points in Step 6 are close together on the collapse fragility curve or the IM levels do
not have significant contributions to c, conducting RHAs at a third intensity and
updating the collapse fragility and c estimates based on the results is suggested. A third
point on the collapse fragility curve is also useful for determining how well the fitted
fragility curve captures the data and deciding whether to run more RHAs to improve the
estimate. The third intensity, IM3, should be selected so that it (a) is not near regions of
the collapse fragility curve and the c deaggregation estimated by the first two IM levels;
and (b) has significant contribution to c based on the c deaggregation.

2
Note that the recommendation of a 20% cumulative contribution differs from the 35% cumulative
contribution recommended in the paper by Eads et al. (2013). The reason for this difference is discussed in
Section 4.5.4.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 79

As currently written, the proposed method cannot handle a null probability at IM1
because Step 4 involves updating the collapse fragility curve based on the dispersion.
This situation is, however, very unlikely given that IM1 is aimed at an intensity level
corresponding to 90% of c. Nevertheless, if one were to run into this situation, one could
get around this by running additional analyses to try and compute a non-null probability
or by selecting a higher value of IM1. In the latter case, the null probability can act as a
third point on the collapse fragility curve and can be included in the fitting used in Step 6.
As discussed by Baker (2013), the proper method for fitting points in a collapse fragility
curve is the method of maximum likelihood.

4.5.3 Discussion of method


Several aspects of the proposed method (e.g., the number of intensities, the level of these
intensities, etc.) were obtained through parametric studies, and the steps were refined
through sensitivity studies on the accuracy of the initial estimate. These studies are
discussed in Section 4.5.4. The IM1 and IM2 values are aimed at capturing intensity
intervals that are significant contributors to c and are good intensities for estimating the
collapse fragility curve. The c deaggregation curve is bell shaped with a long tail to the
right, which means that even if initial estimates of the collapse fragility curve are very
rough, there is still a good chance that the IM1 value will have significant contribution to
c. Because of the short tail to the left of the deaggregations peak, IM2 is generally near
the peak of the deaggregation curve. These intensities are also in regions where the
lognormal curve is a good fit to the observed data (i.e., they are typically not at the
extremes of the collapse fragility curve where increased dispersion between observed
data and the fitted fragility curve is common, and therefore running into cases where
none or all of the ground motions cause collapse is very unlikely).

Performing RHAs at only two ground motion intensity levels greatly reduces the
computational effort of the proposed method compared to performing an IDA.
Furthermore, since RHAs are performed at only two intensity levels, a larger number of
ground motions than currently being used can be considered, which leads to a significant
reduction in the statistical (epistemic) uncertainty in the collapse fragility curve and in c.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 80

It should be noted that the proposed method is very general in nature and can be used
with an IM such as Sa(T1), a vector IM such as Sa(T1) and , or any other ground motion
IM. For more about the advantages and use of a vector IM, the reader is referred to Baker
and Cornell (2005b). As will be demonstrated later, the method is not very sensitive to
the initial estimate of the collapse fragility curve used in Step 1, meaning even if one
estimates the median collapse intensity based on a static pushover method in which the
structure forms a different collapse mechanism than it typically forms during RHAs, the
error and uncertainty in the final c estimate are not expected to be significantly affected.

Another advantage of the proposed two-point approach, compared to performing a full


set of IDAs to collapse, is that more attention can be paid to selection and scaling of the
ground motions to the hazard levels associated with the two values of the IM. Ideally, one
should select different ground motions for each of these two intensity levels consistent
with PSHA deaggregation results to identify the magnitude (and therefore also duration),
distance, focal mechanism, and distributions at these intensity levels. This approach
avoids the shortcoming in IDA of using identical ground motions at very different hazard
levels, which disregards differences in spectral shape, duration and other ground motion
characteristics with changes in intensity levels that, as demonstrated by Chandramohan et
al. (2013), can impact the probability of collapse estimate.

4.5.4 Evaluation of the method via a case study


This section describes how certain aspects of the proposed method outlined in Section
4.5.2 were determined, including the number and level of intensities. Two intensities
were used as a base case, and the intensity levels 3 were varied. For each potential set of
intensity levels, the proposed method was executed with a bootstrap sample of N ground
motions to produce an estimate of c. To facilitate comparisons with the case study, the
ground motion sample used to estimate the probability of collapse at IM1 is also used at
IM2. This process was repeated with 1,000 bootstrap samples for a given value of N and
set of intensity levels. Note that only two intensities were used to construct the collapse

3
In the context of the proposed method for estimating c, the term intensity level refers to the percentage
or fraction of cumulative contribution to c that is used to determine the IM value. This meaning of
intensity level is used throughout this discussion.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 81

fragility curve (i.e., a third intensity was not included to improve the c estimate if the
two intensities were close together). Confidence intervals on c were constructed using
the percentile method (Efron and Tibshirani 1993). Each set of candidate intensity levels
was evaluated on three factors: (1) the mean of samples converging to the true c (the c
computed from the IDA results with all 274 ground motions from the case study) as N
increases; (2) the level of uncertainty in the c estimate, which was judged by the width of
the confidence intervals; and (3) the robustness of the method as measured by its ability
to capture (1) and (2) when the initial guess of the collapse fragility curve was bad (the
definition of bad is discussed later).

Figure 4.5 is presented to better understand the relationship between where a given
intensity is located on the collapse fragility (Figure 4.5(a)), c deaggregation (Figure
4.5(b)), and cumulative contribution to c curves (Figure 4.5(c)). Figure 4.5 shows three
collapse fragility curves that have strikingly different median collapse intensities () of
0.5 g, 1.0 g, and 1.5 g as measured by IM = Sa(T1) and dispersions of = 0.4. The
seismic hazard curve described in Section 3.2 and used in the case study of Section 4.4
was used to compute c for all of the fragility curves. To facilitate comparing the shapes
of the c deaggregations, the c deaggregations presented in Figure 4.5(b) are normalized
by the associated value of c so that the area under each deaggregation curve is unity.
Markers indicate the intensities with 90% and 20% cumulative contributions to c, which
are useful reference points discussed later. It should be noted, however, that these are not
the exact points that one estimates in the proposed method because (a) it is very unlikely
that the random sample of ground motions has exactly the specified distribution and (b)
the collapse fragility curve is updated each time RHAs are performed, both of which
affect the c deaggregation and change the relative location of the reference points.
Though not shown here, the relative positions of intensities corresponding to the 90% and
20% cumulative contributions to c on the collapse fragility curves and normalized
deaggregation curves do not change significantly when the dispersion changes to e.g., =
0.3 or = 0.5.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 82

Figure 4.5: Relationship between where a given intensity is located on (a) the collapse
fragility curve; (b) the c deaggregation curve; (c) the cumulative contribution to c curve.

A comparison of the performance of different intensity levels is provided in Figure 4.6,


which shows the error in the mean c of the bootstrap samples as a function of the
intensity levels (i.e., the cumulative contribution to c) used to select IM1 and IM2 for N =
40. The error is computed as the difference between the true c and the mean c of the
bootstrap samples as is expressed as a percentage of the true c. Figure 4.6(b) shows the
results when the initial guess of the collapse fragility curve is = 1 g (measured by IM =
Sa(T1)) and = 0.4. Note that these values are very close to the parameters of the true
fragility curve, = 1.03 g and = 0.39. Figure 4.6(b) shows that for the conditions
previously described the best performance generally results from using (a) a cumulative
contribution of at least 60% for IM1 and a cumulative contribution of approximately 20%
for IM2 or (b) a cumulative contribution between approximately 10% and 20% for IM1
and a cumulative contribution of at least 40% for IM2.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 83

Figure 4.6: Percent error in the mean c of the bootstrap samples with respect to the
true c as a function of the cumulative contribution to c used to select the intensity
levels IM1 and IM2, shown for N = 40 and an initial guess of the collapse fragility curve
of = 0.4 and: (a) = 0.5 g; (b) = 1 g; (c) = 1.5 g. Note: an x indicates that the error
exceeds 50%.

Figure 4.6(a) and Figure 4.6(c) show the error in the mean c of the bootstrap samples for
N = 40 when the initial estimates of median collapse intensity are = 0.5 g and 1.5 g,
respectively, which represent errors of approximately 50% with respect to the true .
The initial estimate of the dispersion is = 0.4 for both cases. The intensity levels with
the lowest errors vary between the cases: the lowest errors for an underestimate of
(Figure 4.6(a)) generally result when using a cumulative contribution between
approximately 50% - 60% for IM1 and at least 30% for IM2, whereas for an overestimate
Chapter 4: Statistical Uncertainty in the Collapse Fragility 84

of (Figure 4.6(c)) the lowest errors generally resulted from using a cumulative
contribution of at least 40% for IM1 and between 20% and 30% for IM2.

It should be noted that for an initial guess of = 0.5 g (Figure 4.6(a)) using an intensity
level with less than 60% cumulative contribution to c to select IM1 generally resulted in a
null (zero) probability of collapse at IM1 for at least 20% of the samples and for nearly all
of the samples when the cumulative contribution was less than 50%. This is not
surprising given that the initial estimate of the median intensity was very low compared
to the true value. Samples with a null probability of collapse at IM1 were not included in
the statistics. Except for the conditions previously described, samples with a null
probability of collapse at IM1 were rare.

Comparing the errors of the plots in Figure 4.6 leads to some general observations about
the best intensity levels to use given that the collapse fragility curve is unknown:
Using a cumulative contribution to c of at least 60% for IM1 is recommended
because the proposed method can give large errors and a non-trivial chance of
obtaining a null probability at IM1 if the median collapse intensity is
underestimated (Figure 4.6(a)).
Cumulative contributions to c of at least 60% for IM1 and between 20% and 25%
for IM2 produce low errors for both a good initial estimate of (Figure 4.6(b)) and
an overestimate of (Figure 4.6(c)).
For the range of intensity levels listed above (i.e., the cumulative contribution to
c is at least 60% for IM1 and between 20% and 25% for IM2), the error in the
mean estimate of c can be very sensitive to the intensity levels chosen when is
underestimated (Figure 4.6(c)), compared to a good initial estimate of (Figure
4.6(b)) and an overestimate of (Figure 4.6(c)). This suggests that inflating ones
best guess of (e.g., by 20%) used in Step 1 of the proposed method may help
minimize the effect of this sensitivity in the case that is underestimated. In the
case that is a good estimate or overestimated, inflating ones guess of is not
expected to significantly affect the error based on Figure 4.6(b) and Figure 4.6(c).
Chapter 4: Statistical Uncertainty in the Collapse Fragility 85

The best set of intensity levels based on Figure 4.6 is a 90% cumulative
contribution to c for IM1 and 20% for IM2. It should be noted that Figure 4.6 only
considers N = 40 and the error in the mean estimate of c. Errors for other ground
motion sample sizes and the level of uncertainty as a function of the intensity
levels are examined next.

Promising sets of intensity levels were investigated in further detail by examining the
convergence in the mean estimate of c and the level of uncertainty in the c estimates as
a function of N. Figure 4.7 shows the error in the mean c of the bootstrap samples as a
function of N for three sets of intensity levels and three initial estimates of . For a good
estimate of (Figure 4.7(b)) and an overestimate of (Figure 4.7(c)), all sets of intensity
levels produce an error in the mean c that is typically less than 5% for N 50. For a large
underestimate of (Figure 4.7(a) where the estimated median collapse capacity is half of
the correct value), the mean c is best estimated using a 90% cumulative contribution to
c for IM1 and a 20% cumulative contribution for IM2 (the line labeled 0.9, 0.2).

Figure 4.7: Error in the mean c of the bootstrap samples with respect to the true c as a
function of N for different intensity levels. The initial guesses of the collapse fragility
curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g; (c) = 1.5 g. Note: legend entries denote
the fraction of cumulative contribution to c that were used to select IM1 and IM2,
respectively. For example, 0.8, 0.2 means that the intensity with an 80% cumulative
contribution to c was selected for IM1 and the intensity with a 20% cumulative
contribution to c was selected for IM2.

Figure 4.8 shows the uncertainty as a function of N for three sets of intensity levels and
three initial guesses of . The level of uncertainty is measured by the width of the 95%
Chapter 4: Statistical Uncertainty in the Collapse Fragility 86

confidence intervals on c, expressed as a percentage of the true c. These results


indicate that he confidence interval widths (Figure 4.8) are much less sensitive to the
intensity levels than the error in the mean c (Figure 4.7), especially when the median
collapse intensity is underestimated (Figure 4.8(a) versus Figure 4.7(a)). Furthermore, the
confidence interval widths are not significantly affected by the initial estimate of the
median collapse intensity (Figure 4.7(a), Figure 4.7(b), and Figure 4.7(c)).

Figure 4.8: Width of the 95% confidence intervals on c, expressed as a percentage of the
true c, as a function of N for different intensity levels. The initial guesses of the
collapse fragility curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g; (c) = 1.5 g. Note: for
description of legend, see caption for Figure 4.7.

To test whether the bootstrap results were influenced by areas of the collapse fragility
curve where the collapse intensities were not a perfect fit to the lognormal distribution
(i.e., if a bias existed in the error or uncertainty associated with a set of intensity levels
due to how well the case study data fit the lognormal curve at the associated intensities),
the simulations were redone using a set of 274 collapse intensities that were exactly
lognormally distributed. This was achieved by taking the inverse of a lognormal CDF
with the true values of and at 274 probabilities that were evenly spaced between 0
and 1. Figure 4.7 was reproduced as Figure 4.9, and the results obtained using the
exactly lognormal collapse intensities were superimposed as dashed lines. Although the
intensity levels shown in Figure 4.9 generally have relatively small errors, use of
exactly lognormal collapse intensities slightly reduces the error in the mean c estimate
for initial guesses of = 1 g (Figure 4.9(b)) and = 1.5 g (Figure 4.9(c)). Though not
shown, a comparison of the width of the confidence intervals revealed that there were no
Chapter 4: Statistical Uncertainty in the Collapse Fragility 87

significant differences whether the case study or exactly lognormal collapse intensities
were used. Because use of exactly lognormal collapse intensities does not significantly
affect the error in the mean c estimate, the level of uncertainty, or the relative
performance of different sets of intensity levels, it was decided that evaluating the
proposed method based on results derived from the case study collapse intensities is
acceptable.

Figure 4.9: Comparison of error in the mean c of the bootstrap samples with respect to
the true c as a function of N for different intensity levels for the case study collapse
intensities (solid lines) and exactly lognormal collapse intensities (dashed lines). The
initial guesses of the collapse fragility curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g; (c)
= 1.5 g. Note: for description of legend, see caption for Figure 4.7.

Considering all the case study results previously discussed (i.e., convergence in the mean
estimate of c, the level of uncertainty, and the performance when the initial guess of the
collapse fragility is bad), it is recommended that IM1 be selected as the intensity with a
90% cumulative contribution to c and IM2 be selected as the intensity with a 20%
cumulative contribution to c.

The proposed method for estimating c is compared to the IDA method, which finds the
collapse intensity of each record via IDA and constructs the collapse fragility curve by
fitting a lognormal distribution to the collapse intensities. The performance of these
methods is evaluated on the required computational effort and the level of uncertainty in
the estimate. Computational effort is quantified by the number of RHAs required to
estimate c. The IDA method used here for finding collapse intensities involves
performing IDAs in 0.1 g increments of Sa(T1) until collapse occurs and then finding the
Chapter 4: Statistical Uncertainty in the Collapse Fragility 88

collapse intensity within a tolerance of 0.01 g using the bisection method. Using this
approach for the case study, an average of 16 RHAs was required to find the collapse
intensity for each ground motion. It is recognized that this approach may not be the most
computationally efficient approach as, for example, the hunt & fill algorithm
(Vamvatsikos and Cornell 2004) can generate an IDA curve with a maximum of 12
RHAs per record; however, as illustrated below the proposed method is significantly
more efficient than either of these two approaches.

Figure 4.10 compares the 95% confidence intervals and sample means of c as a function
of the number of RHAs for the IDA and proposed methods. The results of the proposed
method are shown when the initial guess of the collapse fragility curve is a median
collapse intensity of = 1 g (measured by Sa(T1)) and dispersion of = 0.4. This
represents a very good initial guess compared to the true values of = 1.03 g and =
0.39; however, Figure 4.7 and Figure 4.8 show that these results are comparable to the
results obtained when there is a 50% error in estimating . Figure 4.10 shows that for
the same amount of computational effort (measured by the number of RHAs used to
construct the collapse fragility curve), the proposed method can provide a good estimate
of c with significantly less uncertainty due to ground motion sample size than the IDA
method.

Figure 4.10: 95% confidence intervals on c as a function of the number of RHAs for the
IDA and proposed methods.
Chapter 4: Statistical Uncertainty in the Collapse Fragility 89

Table 4.1 summarizes the computational effort and margins of error in estimating c for
the case study using the proposed and IDA methods. The margin of error is computed as
half the width of the 95% confidence interval normalized by the value of c computed in
the case study of Section 3.5. Also shown in Table 4.1 is the number of ground motions
employed by each of the methods. For the proposed method, the number of RHAs is
twice the N because the collapse fragility curve is estimated at two points. For the IDA
method, the number of RHAs is 16 times N because an average of 16 RHAs were
required to find the collapse intensity. (Note that in Table 4.1 for a given number of
RHAs, the number of ground motions for the IDA method is rounded to the nearest
integer, so the number of RHAs is not always exactly 16 times N.) Table 4.1 shows that
for that same margin of error, the proposed method requires significantly less
computational effort than the IDA method (e.g., for a margin of approximately 50%, the
IDA method requires over three times the number of RHAs as the proposed method).
Table 4.1 also demonstrates that the proposed method can be both less computationally
demanding and more accurate than the IDA method. For example, estimating c using the
IDA method with 25 ground motions requires 400 RHAs and gives a margin of error of
47%, whereas using the proposed method with 100 ground motions requires only 200
RHAs (half the computational effort) for a margin of error of 33% (30% less
uncertainty).

Table 4.1: Computational effort and level of accuracy when estimating c using the
proposed and IDA methods.
Number of
Response History Margin of Error in c Number of Ground Motions (N)
Analyses (RHAs)
Proposed IDA Proposed IDA
60 77% N/A 30 N/A
80 70% 99% 40 5
120 48% 83% 60 8
200 33% 64% 100 13
400 22% 47% 200 25
640 N/A 36% N/A 40
1200 N/A 26% N/A 75
Chapter 4: Statistical Uncertainty in the Collapse Fragility 90

Table 4.1 also includes entries where the margin of error is only reported for one method
due to the sample size being too small (performing an IDAs with N < 5 would lead to
unreasonable errors) or too big (the ground motion set is limited to 274 ground motions).
These entries were provided out of interest (e.g., one might be interested in the margin of
error when using 40 ground motions with the IDA method) and to facilitate comparisons
between the proposed and IDA methods if one wants to assume that the average number
of RHAs to find the collapse intensity of each ground motion by IDA is a number other
than 16 (e.g., if one wants to compare the methods assuming that an average of 10 RHAs
is sufficient to find the collapse intensity).

The performance of the proposed method was also investigated when using three
intensity levels to estimate the collapse fragility curve. For this investigation the third
intensity level was taken as the mean value of the intensities associated with IM1 and IM2.
When implementing the proposed method in practice, the third intensity level should be
chosen as discussed in Section 4.5.2 (i.e., so that it is not near regions of the collapse
fragility curve estimated by IM1 and IM2 and that it has a significant contribution to c
based on the deaggregation). While it is possible to implement these selection constraints
in a computer code (e.g., by defining a number of conditions and setting tolerances on the
minimum difference between the intensity values), this was not done. It is recognized that
taking the third intensity as the mean of the other two is not an intelligent choice in some
situations (e.g., if the first two intensities are very close to each other) so the results
should be interpreted with this in mind. The best-performing set of intensity levels were
again 90% and 20% cumulative contributions to c for IM1 and IM2, respectively, so
comparisons are shown for only this set of intensity levels. Figure 4.11 compares the
error in the mean c estimate when using two or three intensity levels to estimate the
collapse fragility curve as a function of the computational effort, which is measured by
the number of RHAs. Figure 4.11 shows that for the same computational effort,
estimating the collapse fragility at three intensities as opposed to two can reduce the error
in the mean c estimate but only for certain conditions related to the initial guess of the
collapse fragility curve and number of RHAs. For example, when the initial guess of the
collapse fragility curve is = 1.5 g and = 0.4 (Figure 4.11(c)), estimating the collapse
Chapter 4: Statistical Uncertainty in the Collapse Fragility 91

fragility curve at three intensities as opposed to two produces less error in the mean c
estimate when at least 100 RHAs are used (i.e., at least 50 ground motions used to
estimate P(C|IM) when using two intensity levels and at least 33 ground motions used to
estimate P(C|IM) when using three intensity levels). Cases in which estimating the
collapse fragility curve at three intensities produces more error in the mean c estimate
than estimating at only two intensities for the same computational effort typically occur
when the initial estimate of the collapse fragility curve is very rough and less than 33
ground motions are used to estimate P(C|IM) (Figure 4.11(a) and Figure 4.11(c)).

Figure 4.11: Comparison of error in the mean c of the bootstrap samples with respect to
the true c as a function of the total number of response history analyses used to
estimate the collapse fragility curve. Results are shown when using two or three intensity
levels to estimate the collapse fragility. Intensities with 90% and 20% cumulative
contributions to c are used for IM1 and IM2, respectively. When a third intensity is used,
it is selected as the mean of IM1 and IM2. The initial guesses of the collapse fragility
curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g; (c) = 1.5 g.

Figure 4.12 compares the width of the 95% confidence intervals on c, expressed as a
percentage of the true c, as a function of the total number of RHAs used to estimate
the collapse fragility curve. Figure 4.12 shows that, for the same number of RHAs, using
two intensity levels instead of three generally reduces the uncertainty in the c estimate,
although the reduction can be small, for underestimates (Figure 4.12(a)), good estimates
(Figure 4.12(b)), and overestimates (Figure 4.12(c)) of the initial guess of the collapse
fragility curve. The previous results demonstrate that for the same amount of
computational effort, using two or three intensities to estimate the collapse fragility curve
can give good results in terms of the error in the mean c estimate and level of uncertainty
Chapter 4: Statistical Uncertainty in the Collapse Fragility 92

when IM1 and IM2 are selected as the intensities with a 90% and 20%, respectively,
cumulative contribution to c. When using less than 120 total RHAs, case study results
suggest it is better to use two intensity levels opposed to three. Case study results also
show a trade off when using more than 120 RHAs as using two intensities will generally
result in slightly less uncertainty but slightly more error in the mean c estimate. Another
factor to consider when using a third intensity is the potential effort involved in selecting
appropriate ground motions for a third intensity level.

Figure 4.12: Width of the 95% confidence intervals on c, expressed as a percentage of


the true c, as a function of the total number of response history analyses used to
estimate the collapse fragility curve. Results are shown when using two or three intensity
levels to estimate the collapse fragility. Intensities with 90% and 20% cumulative
contributions to c are used for IM1 and IM2, respectively. When a third intensity is used,
it is selected as the mean of IM1 and IM2. The initial guesses of the collapse fragility
curve are = 0.4 and: (a) = 0.5 g; (b) = 1 g; (c) = 1.5 g.

An interesting issue is how the proposed method performs for a different seismic hazard
curve and/or collapse fragility curve, especially with respect to the recommended
intensity levels. As mentioned earlier, the paper by Eads et al. (2013) recommends using
the intensity associated with a 35% cumulative contribution to c for IM2 while the
proposed method in Section 4.5.2 recommends using 20% instead of 35%. These
recommendations were both calibrated via the same case study (i.e., the same structure
and the same site); however, the seismic hazard curves and the damping used in the
structural model were defined differently, which led to differences in the collapse risk
(and associated c deaggregations and cumulative contributions). A comparison of the
collapse risk results for the case study used by Eads et al. (2013) and the case study
Chapter 4: Statistical Uncertainty in the Collapse Fragility 93

presented in Section 3.5 is shown in Figure 4.13. The c deaggregation in Figure 4.13(c)
also includes markers at the intensity with a 35% cumulative contribution to c for the
Eads et al. (2013) case study and at the intensity with a 20% cumulative contribution to c
for the case study presented in Section 3.5. For both case studies these intensities are at or
very near the peak of the associated c deaggregations. This suggests that choosing the
intensity at the peak of the c deaggregation may be a good choice for IM2. This was
investigated for the case study (i.e., choosing IM2 as the intensity associated with the
peak of the c deaggregation in Step 4 of the proposed method), but the results were not
nearly as good. This is possibly because the peak of the deaggregation estimated in Step 4
is not close enough to the peak of the true deaggregation, particularly if the initial
estimate of the collapse fragility curve is very rough.

Figure 4.13: Collapse risk comparison of case study results from Section 3.5 and those
presented by Eads et al. (2013) for: (a) seismic hazard curve; (b) collapse fragility curve;
(c) c deaggregation.

The results of the Eads et al. (2013) case study are used to compare the performances of
the proposed method for two different intensity levels of IM2: the 20% cumulative
Chapter 4: Statistical Uncertainty in the Collapse Fragility 94

contribution recommended based on the case study in Section 3.5 and the 35%
cumulative contribution recommended based on the Eads et al. (2013) case study. A 90%
cumulative contribution to c is used for IM1 because this was the recommended level for
both case studies. Figure 4.14 compares the error in the mean c of the bootstrap samples
for the different IM2 selections. As expected, better results are produced when IM2 is
associated with a 35% cumulative contribution to c. Though the error is generally on the
order of 10% for N > 75 when IM2 is associated with a 20% cumulative contribution to c
(except for Figure 4.14(a) where the error is approximately 5% for all N shown), this is
more than the error that this set of intensity levels produced for the case study of Section
3.5 (see Figure 4.7). This suggests that the best-performing set of intensity levels is
specific to the collapse fragility and seismic hazard curve being considered.

Figure 4.14: Comparison of error in the mean c of the bootstrap samples with respect to
the true c as a function of N for different intensity levels using the Eads et al. (2013)
case study results. The initial guesses of the collapse fragility curve are = 0.4 and: (a)
= 0.5 g; (b) = 1 g; (c) = 1.5 g. Note: for description of legend, see Figure 4.7.

The results presented in Baker (2013) regarding the intensity levels that produced the
smallest standard deviation in c also support the idea that the best-performing set of
intensity levels is specific to the collapse fragility and seismic hazard curve considered.
Using three different seismic hazard curves with the same collapse fragility curve and
N = 40, Baker found a different set of optimum intensity levels for each case: 80% and
15%, 70% and 10%, and 90% and 20%, respectively, cumulative contributions to c. It
should be noted that Bakers results were not obtained using the proposed method: they
were based on simulating 1,000 sets of data from a known distribution of the collapse
fragility curve, estimating the collapse fragility curve at two intensity levels based on the
Chapter 4: Statistical Uncertainty in the Collapse Fragility 95

simulated data, and calculating c from the estimated fragility. Further research is
warranted regarding general (i.e., not specific to a case study) recommendations for
intensity levels; however, limited results suggest that cumulative contributions of at least
70% for IM1 and less than or equal to 35% for IM2 should be used.

4.6 Summary
This chapter dealt with statistical uncertainty in the collapse fragility and c, specifically
the uncertainty associated with using a small number of ground motions to estimate these
quantities. Different theoretical confidence intervals for the probability of collapse
conditioned on ground motion intensity that depend on the number of ground motions
used were discussed. Based on the recommendations of Agresti and Coull (1998), the
Wilson score confidence intervals are used because they are appropriate for all sample
sizes and estimated probabilities, and they have appropriate coverage probabilities.
Bootstrap techniques were used to quantify the uncertainty in the collapse fragility and c
as a function of the number of ground motions for the case study presented in Chapter 3.

A method for estimating c based on estimating the collapse fragility curve at only two
intensities was proposed. The underlying theory is that for equivalent computational
effort a better estimate of the collapse fragility can be obtained by using more ground
motions at fewer intensities as opposed to using fewer ground motions at more intensity
levels. The proposed method showed promising results based on a case study that
compares the computational effort in obtaining c and level of accuracy and uncertainty
in the estimate for this method and an IDA method. It was demonstrated that the
proposed method can significantly increase the reliability of the computed c while at the
same time significantly reducing the computational effort compared to the IDA method.

The proposed method estimates the collapse fragility curve at two intensity levels based
on the results of RHAs. The intensity levels are determined by their cumulative
contributions to c based on a deaggregation of c by intensity, which is updated each
time a point on the collapse fragility curve is estimated. The recommended cumulative
contribution levels for the case study are 90% for the first intensity and 20% for the
Chapter 4: Statistical Uncertainty in the Collapse Fragility 96

second intensity. These levels were chosen by comparing the performance of different
sets of intensity levels based on their convergence in the mean estimate of c to the true
value of c and the level of uncertainty in c estimate for both good and bad initial
estimates of the collapse fragility curve. It was found that the level of uncertainty in the c
estimate was: (1) not significantly affected by the initial estimate of the collapse fragility
curve; and (2) much less sensitive to the set of intensity levels used than the error in the
mean c estimate.

Limited studies suggest that the best set of intensity levels to use may depend on the
collapse fragility curve and seismic hazard curve being considered, so further
investigation in this area is warranted. However, the results suggest that intensities with
cumulative contributions of at least 70% for IM1 and less than or equal to 35% for IM2
should be used.
Chapter 5

Spectral Shape and Structural Response

5.1 Overview
This chapter focuses on spectral shape and structural response with a particular interest in
identifying what makes a particular record damaging to a particular structure. This issue
is studied from different perspectives including a numerical approach that looks at
individual terms in the equation of motion to identify conditions likely to cause large
changes in displacement and an approach based on the response of a single-degree-of-
freedom (SDOF) system to a half-sine acceleration pulse for different pulse durations,
areas under the pulse, and initial conditions. Formation of spectral shape including the
presence of local peaks and valleys and the slope of the spectrum in specific period
ranges is also discussed. A spectral shape typical of damaging records is identified, and a
metric for quantifying the spectral shape of a record called SaRatio is discussed. It is
shown that SaRatio is a better predictor of collapse intensity than other spectral shape
metrics currently used, and methods for incorporating SaRatio in collapse risk
assessments are discussed.

5.2 Literature review


Ground motion records are commonly characterized by their response spectra, as the
response spectrum of a particular record describes the maximum acceleration, velocity, or
displacement response of elastic SDOF systems as a function of the fundamental period
and level of damping. Prediction of the response spectrum of a record has been studied
extensively, and a review of ground motion prediction equations (GMPEs) developed
between 1964 and 2010 that provide the expected value and standard deviation of a
spectral ordinate at a given period conditioned on the magnitude, distance, fault
mechanism, and site conditions is provided by Douglas (2011). Since publication of the
report by Douglas, several new or updated GMPEs have been developed including, for
example, those for shallow crustal earthquakes in the western US as part of the NGA-

97
Chapter 5: Spectral Shape and Structural Response 98

West2 project (PEER 2013), for crustal earthquakes in Europe and the Middle East
(Akkar et al. 2013), and for subduction zone earthquakes (Gregor et al. 2013).

While response spectra can be useful, they describe only the peak responses of elastic
SDOF systems and do not provide information about the number or sequence of large
displacement cycles a system experiences during a ground motion, which can
significantly affect the inelastic response. As noted by Bertero et al. (1976), the types of
excitations taht [sic] induce the maximum response in elastic and inelastic systems are
fundamentally different, and thus the elastic response spectrum is insufficient for
predicting the maximum inelastic dynamic response. Two records with similar spectra
may produce the same elastic response in a system, but one may cause significantly more
damage when inelastic action is considered.

Some of the first major studies to identify what features of a ground motion make it
particularly damaging were conducted by Bertero et al. (1976; 1978). In the latter study
nonlinear analytical models using near-fault records derived from acceleration recordings
during the 1971 San Fernando, California earthquake were used to show that the damage
observed in structures was primarily the result of severe, long duration acceleration
pulses in these records. In a later study by Anderson and Bertero (1987), the authors
concluded that the area under the acceleration pulse, which they refer to as the
incremental velocity, and not necessarily the amplitude of the pulse, is what makes a
record particularly damaging. They noted that records can be particularly damaging when
they contain a pulse that is long compared to the period of the system and has an effective
average acceleration greater than the systems seismic resistance coefficient. The authors
also noted that long duration acceleration pulses are not unique to near-fault records and
that the nonlinear structural response is sensitive to the duration of the pulse relative to
the period of the structure. These findings are consistent with earlier work (Jacobsen and
Ayre 1952) that studied the effects of the pulse area, shape, and duration on the elastic
response of SDOF systems subjected to a single acceleration pulse.
Chapter 5: Spectral Shape and Structural Response 99

Since these studies significant research (e.g., Baez and Miranda 2000; Alavi and
Krawinkler 2001; Cuesta and Aschheim 2001; Menun and Fu 2002; Mavroeidis and
Papageorgiou 2003; Ruiz-Garca and Miranda 2005; Akkar and zen 2005; Baker
2007b; Baker and Cornell 2008) has been conducted into identifying near-fault, pulse-
like records and quantifying their potentially damaging effects on the structure,
particularly because measuring the intensity of the ground motion using the elastic
spectral acceleration at the structures fundamental period, as is commonly done, or in
combination with the spectral shape parameter often fails to accurately predict the
nonlinear structural response for these type of ground motions (Baker and Cornell 2008).
The focus of previous research has largely been on pulses in the ground velocity history,
but notable exceptions include the work of Sucuoglu et al. (1999) and Makris and Black
(2004) who, like Bertero, demonstrate that structural damage is more closely related to
acceleration pulses than velocity pulses. As noted by Sucuoglu et al. and Makris and
Black, the peak ground velocity (PGV) (the maximum value of the velocity pulse) can
properly indicate the damage potential of a record when this value is derived from a
dominant acceleration pulse; however, the PGV can also result from a series of short
duration acceleration pulses and may result in less damaging records. In particular,
Makris and Black showed that for structures with first-mode periods less than 4 s, near-
fault records in which the PGV is derived from a dominant acceleration pulse are
generally more damaging than those in which the PGV is the result of a series of short
duration pulses.

Many studies have found that incorporating information about the spectral shape (i.e.,
information beyond the ordinate at the fundamental period) or peak ground responses
improves the prediction of structural response. The studies discussed herein will focus on
the prediction of displacement-based responses and collapse as Taghavi and Miranda
(2006, Chapter 9) and more recently Bradley et al. (2010a) showed that approaches to
estimate deformation parameters often are not very good to estimate acceleration
demands and vice-versa. Some early studies used the relationship between PGV and peak
ground acceleration (PGA) as a rough approximation of spectral shape. The significance
of the PGV/PGA ratio was first noted by Newmark (1973) who found this ratio to vary
Chapter 5: Spectral Shape and Structural Response 100

according to the site conditions of the recording station. This was later independently
verified by Seed et al. (1976). More recently, Zhu et al. (1988) and Tso et al. (1992)
showed that the ratio of PGA/PGV provides information on the frequency content of the
ground motion and hence provides a rough measure of spectral shape. In particular, Zhu
et al. found that the PGA/PGV ratio of ground motions could strongly influence the
displacement ductility demands on bilinear SDOF systems when records were scaled to a
common value of PGA. They found that the effect of this ratio on strength demands was
significantly reduced for systems with periods greater than 0.5 s when the records were
scaled to a common value of PGV; however, the same was not found for shorter period
systems.

Kennedy et al. (1984) considered the average spectral value between the fundamental
period of the structure and a lengthened period and noted that the relationship between
the average spectral value and the value of Sa(T1) influenced the nonlinear response of
the structure. In particular, they noted that when the average spectral value was greater
than Sa(T1), the ground motion would require less scaling beyond the yield level to
produce a certain ductility compared to a ground motion where the average spectral value
was less than Sa(T1). Sewell et al. (1996) noted that the damage potential of records was
related to the spectral shape (i.e., the spectral breadth, slope, etc. obtained from spectral
amplitudes at several frequencies over the frequency range of interest), and, for records
scaled to the same Sa(T1), the slope of the spectrum at T1 was related to the damage
potential of the record.

Ordaz and Prez-Rocha (1998) incorporated information about spectral shape into their
expression for estimating strength-reduction factors for SDOF systems by using the ratio
of the spectral displacement at the fundamental period (Sd(T1)) to the peak ground
displacement. They found that their expression was at least as accurate as existing
expressions with the additional benefit of being applicable to both firm and soft soil sites.
Akkar and Miranda (2004) found that inelastic displacement ratios (the ratio of peak
inelastic to elastic displacement) of short-period SDOFs were sensitive to the ratio of
spectral velocity at the fundamental period to the PGV, particularly for weak structures
Chapter 5: Spectral Shape and Structural Response 101

with lateral strength ratios greater than four. By incorporating this spectral ratio into the
prediction of inelastic displacement ratios (and therefore peak deformations of short-
period inelastic systems), they were generally able to reduce dispersion in the results by
at least 50%.

Shome and Cornell (1999) found that the dispersion in maximum story drift was
substantially reduced when the response was conditioned on the spectral acceleration at
the second mode period (Sa(T2)) in addition to Sa(T1). Furthermore, they found that the
probability of collapse conditioned on Sa(T1) alone could be significantly different than
when the probability was conditioned on Sa(T2) or Sa(T3) given Sa(T1).

Carballo and Cornell (2000, Chapter 3) investigated various "form functions" to capture
global and local spectral shapes. They defined the global shape ratio as the ratio of the
moving window average of spectral ordinates centered around a softened effective
frequency (lengthened effective period) to the moving window average of spectral
ordinates centered around the fundamental frequency of the structure, which was then
normalized by the ratio of the median value of the spectral ordinate at the softened
effective frequency to the median spectral ordinate at the fundamental frequency, where
the median values are determined from a set of ground motions. The local shape ratio RLS
was defined as the ratio of the spectral ordinate at the fundamental period of vibration of
the structure to the moving window average of spectral ordinates at the fundamental
period. The authors noted that "a value of RLS considerably higher than unity most likely
determines the presence of a local peak . . . and a value lower than unity is likely related
to the presence of a local valley. They computed the response spectra using 100
equally spaced frequencies in logarithmic space ranging from 0.2 Hz to 25 Hz, and the
window size used to compute the moving average consisted of four frequencies. The
authors concluded that the global shape ratio was more important than the local shape
ratio in predicting the nonlinear behavior of structures.

A common method of incorporating spectral shape is to use the ratio of spectral


acceleration at a given period to the spectral acceleration at the fundamental period
Chapter 5: Spectral Shape and Structural Response 102

(Sa(T1)). The following provides a brief overview of some studies incorporating spectral
shape in this manner to improve the estimation of peak interstory drift ratios of nonlinear
systems.

Cordova et al. (2000) used a parameter called RSa that is defined as the ratio of spectral
acceleration at an elongated period (they recommended 2T1) to Sa(T1). They combine
this parameter with Sa(T1) to form a scalar intensity measure and find a significant
reduction in the dispersion of maximum interstory drift ratios conditioned on intensity.
(Note: discussion about scalar intensity measures based on spectral shape is provided in
Chapter 6. The focus of this chapter is on spectral shape measures that when used with
Sa(T1) (or any other spectral ordinate) provide a better estimation of peak structural
response than Sa(T1) alone. This study was referenced because it was one of the earliest
studies to formally define this spectral ratio.)

Jalayer (2003) found that the response of a 0.1 s SDOF is dependent on the spectral shape
ratio when a period of 2T1 or 3T1 is used. She also found that for a 20-story building
(which had a fundamental period of T1 = 3.96 s), the response was dependent on the
spectral shape factor when using a period close to the second mode period (T2 = 1.35 s).
The responses of these structures were typically limited to low levels of nonlinearity.

Baker and Cornell (2008) used a spectral shape ratio as part of a vector intensity measure
to predict the nonlinear structural response to pulse-like ground motions. After examining
periods less than and greater than T1, they found that using a given period of 2T1 was
effective for moderate to high levels of nonlinearity while using a given period of T1/3
was better for low levels of nonlinearity.

More recently, Bojrquez and Iervolino (2011) measured spectral shape via a parameter
called Np that is defined as the ratio of the average (geometric mean) value of spectral
accelerations between T1 and some period greater than T1 to the value of Sa(T1). They
found that averaging over a period range of T1 to 2T1 generally produced the best results
and that use of Np significantly improved prediction of maximum interstory drift ratio,
Chapter 5: Spectral Shape and Structural Response 103

hysteretic energy demands, and damage over a wide range of nonlinearity levels
compared to using Sa(T1) alone. Furthermore, they found that averaging over a period
range as opposed to using the spectral value at a single ordinate generally provided better
results (less dispersion) in structural response estimates, especially for narrow-band
ground motions, and reduced the sensitivity to the choice of extended period (i.e., the
maximum period used for averaging or the period used for the single spectral ordinate).

Spectral shape is an important consideration when choosing which ground motions to use
for response history analysis. As demonstrated by previous studies the spectral shape of a
record can significantly influence the nonlinear response of a structure; however, the
spectral shape can also influence the linear response of a multiple degree of freedom
structure when using records scaled to a common Sa(T1) because of higher-mode effects.
Design codes often have provisions about the spectral shapes of the records used in
response history analysis. For example, ASCE 7-10 (2010) requires that the average
spectrum of records used in two-dimensional response history analysis cannot be less
than the design spectrum between a period range of 0.2T1 and 1.5T1. Recognizing that
spectral shape can play an important role in nonlinear structural response, the Design
Ground Motion Library (Power et al. 2007), which contains acceleration histories
suitable for response history analysis of various structures in California, evaluates the
spectral shape of a potential record on both (1) the mean squared error of spectral
ordinates over a period band with respect to a target spectrum and (2) the slope of the
records spectrum over the period band versus the slope of the target spectrum over the
same period band.

An important and relevant study conducted by Baker and Cornell (2006a) showed that
the structural response, including the probability of collapse, can be very sensitive to the
method used to select records when Sa(T1) is used to measure the ground motion
intensity. In this study, they selected records based on four different methods: (1)
arbitrarily (i.e., random selection without attempting to match any specific record
properties); (2) based on magnitude and distance; (3) based only on epsilon () at T1; and
(4) based on a spectral shape similar to the conditional mean spectrum, which uses
Chapter 5: Spectral Shape and Structural Response 104

correlations of spectral values at different periods, using the mean magnitude, distance,
and values from probabilistic seismic hazard deaggregation. The spectral shape
parameter measures the difference in the spectral acceleration of a record at a given
period and the spectral acceleration predicted by a GMPE. More information on this
parameter is provided in Section 5.6. Baker and Cornell found that the first two methods
offered biased structural response estimates because the results were dependent on the
values of the records used; however, the latter two methods, which explicitly considered
in record selection, produced unbiased estimates of structural response. This
demonstrated that can be a particularly important proxy to spectral shape. Other
studies regarding the effect of on collapse intensity and how to adjust the collapse
fragility curve to account for this effect include those by Zareian and Krawinkler (2009)
and Haselton et al. (2011).

Another proxy to spectral shape includes the eta () parameter used by Mousavi et al.
(2011), which is a linear combination of the based on spectral acceleration and the
based on PGV. More information on this study is provided in Section 5.6, but the authors
demonstrated that this parameter is even more effective than (based on spectral
acceleration alone) at predicting the collapse intensity of structures.

5.3 What causes large displacement increments?


Damage often occurs as a result of large displacement excursions, so as a first step in
understanding what makes a particular ground motion damaging to a particular system it
can be helpful to understand what causes large displacement excursions. To accomplish
this, individual terms in the equation of motion are examined to determine what
combinations of structural and ground motion properties lead to large displacement
excursions. The equation of motion is first written in its standard form as a function of
time and is then transformed to incremental form and rearranged to solve for the change
in displacement over an increment of time. Equation 5.1 presents the equation of motion
written as a function of time t

mu(t ) + cu (t ) + k (t )u (t ) = mug (t ) (5.1)


Chapter 5: Spectral Shape and Structural Response 105

where m is the mass of the system, c is the damping coefficient, k is the lateral stiffness,
ug is the ground acceleration, and u , u , and u are the systems relative acceleration,

relative velocity, and relative displacement, respectively. Writing the equation in


incremental form yields

mui + cu i + k i u i = mug ,i (5.2)

where in front of a parameter denotes the change in value from time step i to time step
i+1. Rearranging Equation 5.2 to solve for the change in displacement over a given time
increment t gives

mug ,i mui cu i
u i = (5.3)
ki

Using Newmarks method (1959), a single-step integration method for solving the
equation of motion, Equation 5.3 can be rewritten as

mug ,i + au i + bui
u i = (5.4)
k i

where

c m (5.5)
ki = k i + +
t (t )2

m c (5.6)
a= +
t

m
b= + ct 1 (5.7)
2 2
Chapter 5: Spectral Shape and Structural Response 106

and and are constants of Newmarks method that describe how the acceleration is
assumed to vary over the time increment t. To accurately capture the ground
acceleration and the response of the system, t must be small. For linear elastic systems,
using the t of the ground motion recording is generally adequate, but analysis of highly
nonlinear systems can require t values on the order of 10-4 s.

To find the change in displacement over a time period Nt, Equation 5.4 can be
rewritten as

N N mu
g ,i + au i + bui
u N +1 u1 = u i = (5.8)
i =1 i =1 ki

From Equation 5.1 the relative acceleration at time step i is given by

mug ,i cu i k i u i
ui = (5.9)
m

Using Newmarks average acceleration method where = and = and substituting


Equation 5.9 into Equation 5.8 yields

N N ( )
m ug ,i +1 + ug ,i +
4m
u i 2k i u i
u N +1 u1 = u i = t (5.10)
i =1 i =1 ki

which gives the change in displacement over a time period Nt (or equivalently from
time 1 to time N+1).

To understand what causes a large displacement increment over a given time period, it
helps to examine the terms in Equation 5.10. Equation 5.10 shows that the displacement
increment is a function of the mass, damping (via k ), and stiffness of the system as well
as the ground acceleration and the systems velocity and displacement conditions during
the time period.
Chapter 5: Spectral Shape and Structural Response 107

As an example, consider what causes a large displacement increment in a linear elastic


system as it moves from zero displacement to a positive displacement peak, which
typically takes T/4 seconds, where T is the period of the system. For a linear elastic
system, k and k are positive constants and Equation 5.10 simplifies to

N N N
(ug ,i +1 + ug ,i ) + t ui 2k ui
4m
N
m
u N +1 u1 = ui = i =1 i =1 i =1 (5.11)
i =1 k

Because the displacement increment is positive for this example, terms in Equation 5.11
with positive values add to the response, while terms with negative values decrease the
response. The velocities and displacements of the system will, by definition, be positive
because the system is moving in the positive displacement direction. Due to the signs
associated with these terms in Equation 5.11, the velocity term will be positive and the
displacement term will be negative. The velocity decreases to zero as the system reaches
a displacement peak, so a large positive initial velocity (the velocity of the system as it
passes through zero displacement) can be a contributing factor to a large peak
displacement value.

What really causes a large displacement increment, though, is the ground acceleration,
specifically the summation of the ground acceleration values over the time period. This
can be thought of as the area under the ground acceleration, which is the incremental
velocity identified by Bertero and colleagues (1976; 1978). Strictly speaking, the
summation of ground accelerations term as it appears in Equations 5.10 and 5.11 (without
the mass) would need to be multiplied by a factor of t/2 to equal the area under the
ground acceleration using the midpoint integration rule, but for the purpose of this paper
it is useful to think of this term as the area under the ground acceleration because the two
are directly proportional. A large area under the ground acceleration over a given time
period can occur when the acceleration values are very large over a short period of time
(relative to the time period of interest), but this is not very common because these
acceleration values need to be much larger than neighboring values. Most often, large
Chapter 5: Spectral Shape and Structural Response 108

areas occur when the ground acceleration sustains a positive or negative value over some
duration (i.e., a long duration pulse) that is significant with respect to the time period of
interest (i.e., a long duration relative to the period of vibration of the system). For the
example of the linear elastic system moving from zero to a positive displacement peak,
the displacement is maximized when the ground acceleration is negative with a large area
over T/4 and the system has a large positive velocity (momentum) as it passes through
zero displacement.

In general, the largest displacement increments usually occur when there is significant
area underneath the ground acceleration and the ground acceleration acts in the opposite
direction the system is moving (i.e., when the ground acceleration and velocity of the
system have different signs). This leads to a large value of the numerator in Equation
5.10. (Note that here the terms small and large are used to refer to the amplitude or
absolute value and not the sign.) The effects of the area under the ground acceleration and
the systems direction of motion are quantified in Section 5.4. Large displacement
increments in a given system can also be generated when the stiffness values k and k are
relatively small, which can happen when the system is yielding. This causes a small value
of the denominator in Equation 5.10. Obviously the combination of large area under the
ground acceleration, the system moving in the opposite direction of the ground
acceleration, and a small stiffness will also produce a large displacement increment.

5.4 Factors influencing spectral shape


This section examines factors influencing spectral shape including the area and duration
of individual pulses in the ground motion and the displacement and velocity of the system
when pulses begin acting on the system. In this document the term pulse is used to
refer to a segment of the ground acceleration history between zero crossings. The
formation of local spectral features such as peaks and valleys are also examined.

5.4.1 Response to individual pulses


An earthquake ground motion can viewed as a series of individual acceleration pulses.
This section examines how different systems respond to a single half-sine pulse that is
Chapter 5: Spectral Shape and Structural Response 109

derived from a recorded ground motion. Figure 5.1 shows the east-west component of the
ground acceleration recorded at the TCU110 station during the 1999 Chi-Chi, Taiwan
earthquake, which is denoted TCU110E. This recording was obtained from the PEER
NGA database (Chiou et al. 2008). The pulse with the largest incremental velocity (the
area underneath the pulse) is highlighted in a contrasting color. This pulse starts at 43.14
s and has an area of 97.3 cm/s and a duration td of 1.30 s, where td is defined as the time
between zero crossings of the acceleration. Note that td is not equal to the Tp parameter
used by some researchers (e.g., Alavi and Krawinkler 2001; Baker and Cornell 2008) to
identify the period of velocity pulses.

Figure 5.1: TCU110E ground acceleration recording from the 1999 Chi-Chi earthquake.

This pulse is idealized as a half-sine pulse with the same area and duration because
analytical solutions exist for this type of loading and are used later. The original pulse
and the idealized pulse (the equivalent pulse) are shown in Figure 5.2. Other
researchers (e.g., Jacobsen and Ayre 1952; Chopra 2001) have found that the response is
primarily dependent on the area and duration of the pulse while the shape of the pulse
plays a secondary role, especially when the pulse is short with respect to the period of the
system (e.g., td/T < ).
Chapter 5: Spectral Shape and Structural Response 110

Figure 5.2: Acceleration pulse with the largest incremental velocity from the TCU110E
record and an equivalent half-sine pulse with same area and duration.

5.4.1.1 Using at-rest initial conditions


Figure 5.3 shows the linear elastic, 5%-damped displacement response spectrum for the
equivalent half-sine acceleration pulse. To remove the effect of initial conditions, each
system starts with at-rest initial conditions (i.e., zero displacement and zero velocity).
Two displacement spectra are shown: one only considers the peak response while the
pulse acts (the excluding free vibration case), and the other considers that the peak
response can occur at any time, either during or after the pulse. The latter is denoted the
including free vibration case. The two spectra are identical for T 2.64 s
(approximately td/T 0.5) because the peak response occurs while the pulse acts. For T >
2.64 s (approximately td/T < 0.5) the peak response occurs during the free vibration phase
and increases approximately linearly with T. Because an earthquake acceleration pulse
capable of producing any significant structural response would be followed by another
pulse (which would be in the opposite direction but of unknown duration and intensity),
the free vibration phase is not considered herein. All subsequent discussions and spectra
will disregard the free vibration phase after the pulse unless noted otherwise.
Chapter 5: Spectral Shape and Structural Response 111

Figure 5.3: Displacement response spectrum for the equivalent half-sine acceleration
pulse using at-rest initial conditions.

Figure 5.3 shows that the spectral displacement (Sd) values increase sharply for 0.52 s < T
< 2.64 s (0.5 < td/T < 2.5, approximately). The sharp increase occurs because the peak
response of these systems occurs at the first local minima in the displacement response.
For systems where T < 0.52 s (td/T > 2.5), the peak response does not occur until at least
the second local minima. The peak response transitioning from later local minima to the
first local minima is shown in Figure 5.4(a). Systems where T > 2.64 s do not reach a
local minima (i.e., zero velocity) before the pulse ends, so the Sd value is taken as the
displacement at the end of the pulse. The peak response transitioning from the first local
minima to the end of the pulse is shown in Figure 5.4(b).
Chapter 5: Spectral Shape and Structural Response 112

Figure 5.4: Displacement response and occurrence of peak response for selected systems
subjected to a half-sine acceleration pulse: (a) peak response transitioning to first local
minima; (b) peak response transitioning from first local minima to end of pulse. (Note the
different vertical scales.)

As seen in Figure 5.3, the Sd values approach an asymptote as T increases beyond 2.64 s.
The value of the asymptote can be found by solving the equation of motion for a system
subjected to a harmonic load to obtain the displacement response (a readily available
solution), then finding the equation constants using at-rest initial conditions, next writing
the displacement equation at time t = td (which is when systems where T >> td experience
their peak displacement), and finally taking the limit of the displacement as T approaches
infinity. When doing this one finds that for a system with at-rest initial conditions
subjected to a half-sine pulse, Sd approaches a value equal to half the area of the pulse
(Ap) multiplied by the pulse duration td. In other words

Aptd
lim S d (T ) =
T 2 (5.12)

As one might expect, this value is equal to the ground displacement at the end of the
pulse. When T >> td the system barely moves while the pulse acts (it will not move if
T ), and the relative displacement is due to the motion of the ground.
Chapter 5: Spectral Shape and Structural Response 113

The 5%-damped displacement response spectrum shown in Figure 5.3 can be generalized
for a half-sine pulse of arbitrary area Ap and duration td. The generalized spectrum is
shown in Figure 5.5. The vertical axis is the spectral displacement normalized by the
spectral displacement of a system with an infinitely long period of vibration. The
horizontal axis is the period of the system normalized by the pulse duration. This is the
signature displacement spectral shape of a half-sine pulse for systems with at-rest
initial conditions. It increases sharply where the period of the system is between
approximately one and three times the pulse duration (1 < T/td < 3) and begins to flatten
out at larger periods. Knowing these shape features can be useful in deducing properties
about the acceleration input given a particular spectrum. For example, the period range
over which the Sd values sharply increase provides information about the duration of the
pulse. Though this spectrum is specific to half-sine pulses and systems with at-rest initial
conditions, some of its basic features are applicable to spectra from earthquake ground
motions. This will be discussed later.

Figure 5.5: Normalized displacement response spectrum given a half-sine acceleration


pulse and at-rest initial conditions.

To emphasize the effect of the incremental velocity (i.e., the area of the pulse Ap) on the
response, Figure 5.6 shows how Sd increases with increasing incremental velocity. Figure
5.6(a) shows a half-sine acceleration pulse with different pulse amplitudes (p) while
Figure 5.6(b) shows the Sd normalized by the square of the pulse duration td versus the
period of the system normalized by td. For a given system and a given pulse duration (i.e.,
a given T/td), the spectral displacement is linearly related to incremental velocity.
Chapter 5: Spectral Shape and Structural Response 114

Figure 5.6: Effect of half-sine pulse amplitude on displacement demands of SDOFs with
at-rest initial conditions: (a) acceleration pulse; (b) normalized displacement spectrum.

5.4.1.2 Effect of initial conditions


This section examines how the response to the same half-sine acceleration pulse is
affected by the displacement and velocity (d0p and v0p, respectively) conditions of the
system at the beginning of the pulse. Herein the term initial conditions is used to refer
to the state of the system (i.e., its displacement d0p and velocity v0p) at the onset of a
ground motion pulse. A T = 2.5-s system with 5% damping is used for illustration, and
the initial conditions are derived from the systems displacement and velocity conditions
during the TCU110E record shown in Figure 5.1. At 43.14 s when the pulse strikes, the T
= 2.5-s system has a relative displacement of 59.43 cm and a relative velocity of 120.21
cm/s. These values are used as the initial conditions for the equivalent half-sine pulse
shown in Figure 5.2 to study how the signs of the initial conditions affect the response.
The effect of the magnitude of the initial conditions on the response will be discussed
later.

Figure 5.7(a) shows the displacement response of the T = 2.5-s system to the half-sine
pulse shown in Figure 5.7(b) for five different initial conditions. Except for the case with
at-rest initial conditions, the magnitudes of the initial conditions are the same but the
signs vary in each case. Figure 5.7(a) demonstrates that the initial conditions can have a
significant effect on the response. The maximum displacements range from 28 cm (at-rest
initial conditions) to 89 cm (both systems with positive d0p). The two systems with
positive d0p conditions lead to the greatest maximum displacements because these
Chapter 5: Spectral Shape and Structural Response 115

systems move in the same direction as the momentum created by the pulse (i.e., the
systems velocity is opposite the ground acceleration) for the majority of the pulse
duration. This is not surprising given the discussion in the previous section about the
terms in Equation 5.10 and what conditions lead to large displacement increments. In this
particular case, the sign of v0p does not affect the maximum response of the systems with
positive d0p, but it does affect the time at which the maximum displacement occurs. The
system with positive d0p and positive v0p reaches its peak of 89 cm after the pulse stops. If
only the peak displacement while the pulse acts is considered, the system with positive
d0p and positive v0p has a peak displacement of 83 cm.

Figure 5.7: Response to a half-sine pulse with different initial conditions: (a)
displacement response of a 2.5-s system; (b) half-sine acceleration pulse.

In this example the sign of v0p has a larger effect on the maximum response of systems
starting with a negative d0p. The system with negative d0p and negative v0p moves in the
same direction as the pulse for only the first 0.29 s, which is when it reaches it maximum
displacement of 77 cm. The system with negative d0p and positive v0p moves in the
opposite direction of the pulse momentum for the first 0.89 s. At this time it reaches a
local maximum of 45 cm, but this is smaller than its initial displacement of 59 cm
because the pulse counter-acts the motion of the system. Referring back to Equation 5.10,
it is not surprising that this system has the smallest peak displacement of all the (non-at-
rest) d0p and v0p combinations because the summation of the ground acceleration term and
Chapter 5: Spectral Shape and Structural Response 116

the system velocity term oppose each other over most of the pulses duration, which
results in a smaller displacement increment.

The previous example considered how the peak displacement response to a particular
pulse varied for specific initial conditions: at-rest and for a particular value of d0p and a
particular value of v0p, with variations of the signs of the d0p and v0p values. To
understand in a more general way how the state of the system (the d0p and v0p conditions)
at the onset of the pulse affects the displacement response and, in particular, the peak
displacement demand, Figure 5.8(a)-(e) shows the peak displacement plotted as a
function of the time during the vibration cycle at which the pulse starts for various td/T
ratios. Like the previous example, this figure uses linear elastic, 5%-damped SDOF
systems and a positive-valued, half-sine acceleration pulse. The peak displacement of the
system is taken as the peak response while the pulse acts (i.e., the free vibration portion
of the response is not considered). The peak displacement is normalized by Sd0, the peak
displacement using at-rest initial conditions, so that the figure applies to half-sine pulses
of any area. Values larger than one on the vertical axis represent a measure of
amplification of displacement demands caused by the state of the system at the onset of
the pulse. Although a few systems have values smaller than one (i.e., some initial
conditions cause systems to experience responses smaller than they would experience
with at-rest initial conditions), in most cases initial conditions different than zero result in
an increased response.

The d0p and v0p conditions are derived from harmonic motion, an approximation to how
systems vibrate during earthquakes. The assumed initial displacement d0p of the system
normalized by some displacement amplitude dAmp is shown in Figure 5.8(f) as a function
of t/T, where t is the time at which the pulse arrives and T is the period of vibration of the
system. The v0p condition, normalized by 2dAmp, is also shown in Figure 5.8(f). The
displacement amplitude dAmp describes the peak displacement amplitude at which the
system was vibrating before the pulse hit and is defined with respect to Sd0. As Figure
5.8(a) demonstrates, increasing dAmp generally leads to an increase in peak
Chapter 5: Spectral Shape and Structural Response 117

displacement, meaning that for a given pulse the peak response can be significantly
affected by the displacement amplitude of the cycle preceding the pulse.

Figure 5.8: (a)-(e) Peak displacement as a function of pulse arrival time t for various
amplitudes of initial conditions: (a) td/T = 0.1; (b) td/T = 0.3; (c) td/T = 0.5; (d) td/T = 0.75;
(e) td/T = 1.5; (f) initial conditions of the system as a function of pulse arrival time t.

The time during the displacement cycle at which the pulse hits also affects the response.
As an example, consider Figure 5.8(b) for td/T = 0.3 and dAmp = 2Sd0. The largest
response occurs when the pulse hits at approximately t/T = 0.45. The d0p and v0p
Chapter 5: Spectral Shape and Structural Response 118

conditions at t/T = 0.45 in Figure 5.8(f) show that the displacement is slightly positive
and the velocity is negative when the pulse hits. Assuming the system vibrates
harmonically, the system will continue to move in the negative displacement direction
(i.e., have a negative velocity) until approximately t/T = 0.75, which is also when the
pulse stops (because the duration of the pulse is 0.3T and it starts at t/T = 0.45). This
means that for nearly the entire time the pulse was acting, the system was moving in the
same direction as the momentum imparted by the pulse. This is why the peak response
was larger when the pulse hit at t/T = 0.46 versus t/T = 0.93, the time associated with the
minimum response. When the pulse hits at t/T = 0.93, the displacement is negative but
close to zero and the velocity is positive and close to a local maximum (Figure 5.8(f)). In
this case the motion of the system and the momentum imparted by the pulse oppose each
other for almost the entire duration of the pulse.

The td/T ratio also determines how significantly the timing of the pulse affects the
response. If the pulse is very short with respect to the period of the system, the time at
which the pulse hits (i.e., the state of the system when the pulse hits) can significantly
affect the peak response as illustrated in Figure 5.8(a); however, these types of pulses
typically cause smaller displacements than longer duration pulses. One initial state could
cause the systems motion and the pulses momentum to be in the same direction for the
duration of the pulse whereas another initial state could cause them to oppose each other.
If the pulse is long with respect to the period of the system (e.g., td/T = 1.5), the time at
which the pulse hits has much less of a role because the system will move in the same
direction as the pulses momentum for part of the time and in the opposite direction for
the other part of the time. The effect of td/T ratio with respect to how significantly the
pulse arrival time affects the peak response can be seen by comparing the plots for td/T =
0.1 (Figure 5.8(a)) and td/T = 1.5 (Figure 5.8(e)) for a given displacement amplitude. For
example, when dAmp = 2Sd0 the normalized peak response varies from 0.3 to 3.0 for td/T
= 0.1, whereas it only varies from 2.0 to 2.6 for td/T = 1.5.
Chapter 5: Spectral Shape and Structural Response 119

5.4.2 Response to a ground motion


This section focuses on the peak response to a ground motion and the variation in
response for systems with similar periods and elastic versus inelastic systems. The
formation of spectral features such as peaks and valleys is also discussed.

5.4.2.1 Peak elastic response of SDOF systems with similar periods


The displacement response spectrum for the Chi-Chi ground motion is shown in Figure
5.9(a). Figure 5.9(b) displays the Time at Peak or T@P spectrum, which shows the
time in the ground acceleration history at which each system reaches its peak
displacement (i.e., the time at which the spectral displacement, Sd, or pseudo-spectral
acceleration ordinate occurs). As the T@P spectrum demonstrates, different SDOF
systems reach their peak displacement at different times during the ground motion. The
discontinuities in the T@P spectrum indicate that different portions of the ground motion
control the response of different systems. For example, the smooth Sd peak from T = 2.21
s to 3.10 s (Figure 5.9(a)) occurs because SDOF systems with periods in this range reach
their peak displacements at approximately the same time (at approximately 44 s as seen
in Figure 5.9(b)). Similarly, SDOF systems with periods of vibration between 3.29 s and
3.81 s reach their peak displacement approximately 41 s into the record.

Figure 5.9: TCU110E record: (a) displacement response spectrum; (b) Time at Peak
(T@P) spectrum.
Chapter 5: Spectral Shape and Structural Response 120

It is useful to look at the displacement history of individual systems whose peak


responses occur at approximately the same time to understand why they exhibit different
amplitudes in their peak response, given that their peak response is controlled by the
same portion of the ground motion. For example, why is the T = 2.5-s system at the top of
the spectral peak while the T = 2.21-s and T = 2.8-s SDOF systems experience much
smaller peak displacements even though their peaks occur at approximately the same
time during the record? These three points are marked with an x on the displacement
spectrum in Figure 5.9(a). A relevant portion of the displacement history of these three
systems is shown in Figure 5.10(a) with the corresponding ground acceleration history
shown in Figure 5.10(b). The vertical lines at 43.14 s and 44.44 s denote the beginning
and end, respectively, of the acceleration pulse with the largest incremental velocity in
this record. The three systems reach their peak displacement (denoted by circles in Figure
5.10(a)) in the negative direction because a positive acceleration pulse tends to move the
systems in the negative direction.

A significant reason why the peak of the displacement spectrum occurs around T = 2.5 s
is because this system has very unfavorable initial conditions when the critical
acceleration pulse begins to act on the system. The initial conditions are described here as
unfavorable because they lead to larger displacement demands on the structure. Of all
the elastic systems subjected to this ground motion, systems at the top of the spectral
displacement peak near T = 2.5 s in Figure 5.9(a) represent a worst case scenario in
terms of initial conditions when the acceleration pulse hits, the area of the pulse (the
incremental velocity), and the duration of pulse with respect to the period of the system.
As Figure 5.10(a) shows, the T = 2.5-s system is about to reach a positive displacement
peak when the pulse starts. Once a system reaches a positive displacement peak, it
unloads and starts to move in the negative displacement direction. The acceleration pulse
provides additional momentum for the system to move in the negative displacement
direction, which increases the response. This particular pulse also has the largest
incremental velocity of any pulse in the ground motion, which means it imparts the most
momentum. Finally, the duration of this pulse is 1.3 s, which, when paired with the initial
conditions of the system, means that it acts nearly the entire time that the system is
Chapter 5: Spectral Shape and Structural Response 121

moving in the negative displacement direction (it takes approximately T/2 s to move
between local maxima and minima).

Figure 5.10: TCU110E record: (a) displacement response for systems in the spectral
displacement peak; (b) ground acceleration.

The 2.21-s and 2.8-s SDOF systems experience smaller peak displacements because the
initial conditions when the pulse starts to act are more favorable compared to the 2.5-s
system. The T = 2.8-s system is at the toe of the spectral peak because the combination of
initial conditions when the critical pulse begins to act and pulse duration relative to the
period of vibration of the system is more benign than for the T = 2.5-s system. Figure
5.10(a) shows that, time-wise, the 2.8-s system has completed approximately of its
displacement cycle between the local minima at t = 42.36 s and the start of the critical
pulse at t = 43.14 s. The system needs to continue moving in the positive displacement
direction for approximately T/4 (approximately 0.7 s) before it can start moving in the
negative displacement direction, which is the direction of the pulses momentum. The
result is that the motion of the 2.8-s system and the momentum imparted by the pulse
oppose each other longer than they did for the 2.5-s system. In addition to the timing of
the initial conditions, the displacement amplitude of the 2.8-s system in previous cycles is
smaller than the amplitude of the 2.5-s system, which as discussed in Section 5.4.1.2
affects the peak response to a pulse. Finally, the pulse stops acting before the 2.8-s
system finishes moving in the negative displacement direction, so as the system is
moving towards a negative displacement peak, the momentum imparted by the next pulse
is in the positive direction, which decreases the peak response. Though this also occurs
Chapter 5: Spectral Shape and Structural Response 122

for the 2.5-s system, the pulse stops acting much closer to where the 2.5-s system reaches
its negative displacement peak.

Though the T = 2.21-s and 2.8-s systems have approximately the same displacement
amplitudes in previous cycles, the 2.21-s system has a greater peak displacement because
it has less benign initial conditions (it is just starting to move in the negative
displacement direction when the pulse starts) and the pulse acts the entire time the system
is moving in the negative direction. Even though the pulse acts approximately the entire
time the 2.21-s and 2.5-s systems move in the negative displacement direction and both
systems have approximately the same displacement when the pulse starts, the 2.21-s
system does not reach a peak displacement as large as the 2.5-s system because the
former slows down considerably and eventually unloads prior to the end of the pulse.
Additionally, the 2.21-s system has a smaller displacement amplitude in the previous
cycle compared to the 2.5-s system.

5.4.2.2 Formation of spectral peaks and valleys


The previous section examined conditions leading to differences in spectral ordinates
among systems in the peak of a displacement spectrum, namely favorable or
unfavorable combinations of: (a) the state of the system when a pulse hits (i.e., the
displacement and velocity); (b) the area under the acceleration pulse; and (c) the duration
of the pulse with respect to the period of the system. Therefore, spectral shape is a
combination of the characteristics of individual pulses contained in the ground motion
and the initial conditions of the systems when the critical pulses arrive. Smooth spectral
regions indicate that the peak responses of those systems are controlled by the same
portion of the ground motion (i.e., adjacent systems reach their maximum displacements
at approximately the same time). For the example ground motion in Figure 5.9 this occurs
most notably for the smooth spectral peaks from T = 2.21 s to 3.10 s and from T = 3.29 s
to 3.81 s.

Smooth spectral valleys can also occur, but these are less common. Unlike smooth
spectral peaks, smooth (i.e., U-shaped) spectral valleys are formed by benign
Chapter 5: Spectral Shape and Structural Response 123

combinations of system and pulse properties that create reduced responses (e.g., the
system moving in the opposite direction of the pulse momentum). When the spectrum
tends toward a smooth valley, the response to that segment of the ground motion can
become so small that it is no longer the maximum response, and the maximum
displacement is produced by a different portion of the ground motion. When this occurs a
V-shaped valley is often formed in the spectrum, such as the valley at T = 3.29 s in
Figure 5.9. The peak response shifting from one portion of the ground motion to another
is illustrated in Figure 5.11, which shows the displacement history for T = 3.19 s, 3.29 s,
and 3.39 s. The peak displacement of the 3.19-s system occurs just prior to 44 s (at t =
43.75 s). The local peak around this time produces the spectral displacement for systems
from T = 3.11 s to 3.28 s (Figure 5.9). As the period increases from 3.19 s to 3.29 s, the
displacement at the local peak just prior to 44 s decreases while the displacement at the
local peak just after 41 s increases. When the period reaches 3.29 s, the displacement at
the local peak just after 41 s becomes slightly higher (by 1 cm) than that of the local peak
just before 44 s. As the period increases from 3.29 s to 3.39 s, the local peak just after 41
s continues to grow and produces the spectral displacement for systems from T = 3.29 s
to 3.81 s (Figure 5.11 and Figure 5.9).

Figure 5.11: TCU110E displacement response of systems near the spectral displacement
valley at T = 3.29 s, illustrating the peak displacement shifting from one local maxima to
another as the period changes.

Examination of the displacement and T@P spectra in Figure 5.9 shows that the V-
shaped valleys or abrupt changes in the slope of the displacement spectrum, except those
in the very short period range, are created when there is a change in the controlling
portion of the ground motion (i.e., when there are large discontinuities in the T@P
Chapter 5: Spectral Shape and Structural Response 124

spectrum). Systems located where there is a shift in controlling portion of the ground
motion contain at least two local peaks with amplitudes approximately equal to the
spectral displacement (obviously at least one of the peaks exactly equals the spectral
displacement). The two local peaks can be in opposite displacement directions (i.e., one
positive and one negative) or can be in the same direction, such as the 3.29-s system in
Figure 5.11.

5.4.2.3 Elastic versus inelastic response


As discussed in previous sections, peaked regions in the displacement spectrum are often
created by some SDOF systems having more unfavorable initial conditions than others
when hit by long duration acceleration pulses, among other factors. When systems in a
peaked elastic spectral region experience yielding prior to the arrival of the critical
acceleration pulse, the initial conditions for the critical acceleration pulse will be different
and in most cases more benign than those experienced by the elastic system. This is why
systems in a peaked elastic spectral region often have smaller inelastic responses
compared to elastic responses. Conversely, when systems are in the valley of an elastic
spectral region, often because they had favorable initial conditions, inelastic systems with
prior yielding excursions will exhibit different initial conditions that in most cases will
not be as benign as those of the elastic system, and therefore, the inelastic systems will
typically exhibit an increased response compared to the elastic systems. This partially
explains why some researchers (e.g., Baker and Cornell 2006a; Haselton et al. 2011) have
recently found that for the same Sa(T1) value, records with a peaked shape at or near T1
tend to be less damaging than records without a peaked shape at or near T1.

Figure 5.12 compares the elastic and inelastic displacement response spectra of the same
record used in the previous section. The inelastic systems are elastic, perfectly plastic
bilinear systems with R values of 2 and 4, where R is the ratio of the strength required for
elastic behavior to the strength of the actual system. As seen in Figure 5.12, the elastic
spectral peak from T = 2.21 s to 3.10 s disappears in the inelastic spectra.
Chapter 5: Spectral Shape and Structural Response 125

Figure 5.12: Elastic and inelastic displacement response spectra for the TCU110E record.

A selected portion of the elastic and inelastic displacement histories for T = 2.5 s is
shown in Figure 5.13(a) along with the ground acceleration history in Figure 5.13(b). It is
clear that at the start of the pulse slightly after 43 s the initial conditions of the inelastic
systems are different than the unfavorable initial conditions that led to the large peak in
the elastic displacement spectrum at 2.5 s. In particular, Figure 5.13 shows that both
inelastic systems exhibit smaller displacements and velocities at the onset of the critical
acceleration pulse than those in the elastic system. The smaller displacement amplitudes
lead to less potential energy stored as strain energy in the system, which translates to
significantly smaller peak velocities, and as a result the inelastic systems have decreased
peak responses compared to the elastic system. The inelastic systems with R values of 2
and 4 reach peak displacements that are approximately 65% and 60%, respectively, of the
91 cm peak displacement reached by the elastic system.
Chapter 5: Spectral Shape and Structural Response 126

Figure 5.13: TCU110E record: (a) displacement history for T = 2.5 s; (b) ground
acceleration.

5.5 Spectral shape of damaging records


This section focuses on the spectral shape of damaging records and discusses why these
records have distinct spectral shapes and what features in the ground motion give rise to
this shape. Results from the case study of the 4-story office building with steel special
moment-resisting frames described in Chapter 3 are used. Full details of the case study
including structural geometry and design, modeling, site, seismic hazard and the ground
motion set are provided in Chapter 3.

As discussed in Chapter 3, the median collapse intensity of the structure under the set of
274 ground motions is Sd(T1) = 45.3 cm (Sa(T1) = 1.03 g), where T1 = 1.33 s. The
displacement spectra of records producing the 10 lowest and 10 highest collapse
intensities are shown in Figure 5.14, where all records are scaled to Sd(T1) = 10 cm. As
shown in this figure, the spectral shape of the damaging records (those with the lowest
collapse intensities) is significantly different than the spectral shape of the benign
records (those with the highest collapse intensities). With the exception of one record 1,

1
This record is from the 1999 Duzce, Turkey earthquake and was recorded at the Lamont 375 station
(PEER NGA record 1617, first horizontal component). It is not surprising that the spectral shape of this
record does not follow the trend of the other damaging records. This record has unusually high ordinates in
the short period range (e.g., Sd(0.4s) = 31 cm as seen in Figure 5.14) caused by unusual high frequency
content in the record and leads to a collapse mechanism in the 4th (top) story. The overwhelming majority
of the 274 records lead to a collapse mechanism involving the lower three stories.
Chapter 5: Spectral Shape and Structural Response 127

which is shown with a thin red line, the damaging records generally exhibit sharply
increasing spectral ordinates for periods greater than T1. The benign records, on the other
hand, exhibit relatively constant or slightly decreasing spectral ordinates for periods
greater than T1.

Figure 5.14: Displacement spectra of records producing the 10 lowest and 10 highest
collapse intensities, where records are scaled to a common value of Sd(T1).

Why do the damaging records generally exhibit sharply increasing spectral ordinates for
periods greater than T1? A related question is what in the ground motion makes these
records damaging to this particular structure. Recall Figure 5.5 that presents the
normalized displacement response spectrum for a half-sine pulse with at-rest initial
conditions. Figure 5.14 shows that the spectral ordinates sharply increase over a period
range between approximately one and three times the pulse duration (td < T < 3td).
Though this spectrum was for a single pulse and not a ground motion, the individual
pulses that comprise a ground motion give rise to the spectral shape. This will be further
explored by examining selected damaging records in detail.

The first damaging record considered is the north-south component of the ground
acceleration recorded at the TCU063 station during the 1999 Chi-Chi, Taiwan
earthquake, which is denoted TCU063N. The record is shown in Figure 5.15, where all
plots are scaled to Sd(T1) = 10 cm. The 5%-damped displacement and T@P spectra are
shown in Figure 5.15(a) and Figure 5.15(b), respectively. The vertical lines in these
figures separate the spectra into regions based on the pulse that was acting when the
Chapter 5: Spectral Shape and Structural Response 128

system reached its peak (spectral) displacement. All SDOF systems that fall between
adjacent vertical lines reach their peak displacements during the same pulse. Figure
5.15(c) shows the displacement spectra for the full motion and for the pulse in the ground
motion with the largest incremental velocity, which occurs slightly after 40 s into the
record. Two spectra are shown for the pulse: one computed using at-rest initial conditions
and the other computed using the displacement and velocity conditions of the system
from the full ground motion at the start of the pulse. The pulse spectra in Figure 5.15(c)
only consider the response while the pulse acts (i.e., they do not consider peak responses
during free vibration or subsequent parts of the ground motion). The pulse with the
largest incremental velocity, which is shown in Figure 5.15(d) with a relevant portion of
the ground acceleration history, starts at t = 40.37 s and has a duration (td) of 1.66 s.

Figure 5.15: TCU063N record scaled to Sd(T1) = 10 cm: (a) displacement response
spectrum; (b) T@P spectrum; (c) displacement response spectra including spectra for the
pulse with the largest incremental velocity; (d) ground acceleration history.

Similar to the other damaging records in Figure 5.14, this record generally exhibits
sharply increasing Sd ordinates for periods greater than T1 = 1.33 s. The plots comprising
Chapter 5: Spectral Shape and Structural Response 129

Figure 5.15 show that this increasing spectral shape is primarily the result of the pulse
with the largest incremental velocity. The T@P spectrum (Figure 5.15(b)) shows that a
wide range of periods (practically all between 1.1 s and 4.0 s) reach their peak response
during a short window of time (from t = 40.91 s to t = 44.32 s), which includes the largest
incremental velocity pulse and approximately two seconds following the pulse as shown
in Figure 5.15(d). The spectrum of the pulse with at-rest initial conditions (Figure
5.15(c)) forms a base for the spectral shape of the full ground motion. This pulse has a
duration of td = 1.66 s, so the sharp increase in spectral ordinates greater than T1 is
consistent with the sharply increasing spectral ordinates in the range of approximately td
< T < 3td that were observed for a half-sine pulse with at-rest initial conditions (Figure
5.5).

As previously mentioned in Section 5.4.1.2, when the initial conditions from the ground
motion are used to compute the pulse spectrum instead of at-rest initial conditions, the
spectral values typically increase. The initial conditions also result in a more jagged
spectral shape compared to the relatively smooth shape of the spectrum with at-rest initial
conditions (Figure 5.15(c)). Because the pulse with the largest incremental velocity
controls the peak response of most systems with T > 1.68 s, the spectrum of the pulse
with the initial conditions from the ground motion approaches or matches spectrum of the
full ground motion for this period range (Figure 5.15(c)). The only reason that the spectra
do not match between T = 2.49 s and 4 s is because these systems reach their peak
displacement within the half-cycle of displacement following the end of pulse.

The spectral shape of this damaging record, specifically the sharply increasing spectral
displacement ordinates for periods greater than T1, is caused by the pulse with the largest
incremental velocity. This pulse is also the main reason why this ground motion is so
damaging to this particular structure. Figure 5.16(a) shows the story drift ratio (SDR)
history in the first story of the structure for two ground motion intensities: one that causes
significant inelastic behavior (Sa(T1) = 0.5 g) and one that causes collapse (Sa(T1) = 0.6
g). The unscaled ground acceleration history is shown in Figure 5.16(b). As Figure 5.16
shows, the pulse with the largest incremental velocity causes a major inelastic excursion.
Chapter 5: Spectral Shape and Structural Response 130

In addition to its large incremental velocity, this pulse is particularly damaging because
its duration (1.66 s) is close to the first-mode period of the structure (1.33 s), and the
pulse imparts momentum for approximately an entire displacement cycle. It is not
surprising that the structure collapses in the positive displacement direction because the
pulse with the largest incremental velocity has negative acceleration amplitudes and
therefore imparts momentum on the structure in the positive direction.

Figure 5.16: TCU063N record: (a) 1st story drift ratio history; (b) unscaled ground
acceleration history.

Not all damaging records are so dominated by a single pulse. Typically different pulses
control different period ranges of the response spectrum, and significant inelastic
excursions occur at multiple pulses. This is the case for the damaging record shown in
Figure 5.17, which is the north-south component of the ground acceleration recorded at
the TCU055 station during the 1999 Chi-Chi, Taiwan earthquake, which is denoted
TCU055N, and is scaled to Sd(T1) = 10 cm. The 5%-damped displacement spectrum
and the T@P spectrum in Figure 5.17(a) and Figure 5.17(b), respectively, demonstrate
that different pulses control different period ranges. The pulse and full motion spectra in
Figure 5.17(c) show that unlike the previous damaging record a single pulse does form
the basis for the spectral shape for all T > T1, especially between T = 3.2 s and 4 s. The
selected pulse, which is shown in Figure 5.17(d) with the ground acceleration history,
starts at t = 19.26 s and lasts for td = 0.96 s, and it drives the peak elastic response of
SDOF systems between T = 2.24 s and 3.19 s. Quotation marks are used because a pulse
Chapter 5: Spectral Shape and Structural Response 131

was previously defined as a segment of acceleration between zero crossings. The pulse
selected for this ground motion is comprised of 0.57 s of positive acceleration values
followed by 0.04 s of negative acceleration values followed by 0.35 s of positive
acceleration. The two positive segments of the pulse contain the largest and third
largest incremental velocities (measured between zero crossings) of any pulses in the
ground motion. Because the negative segment is short relative to the positive segments,
the three segments are treated as a single pulse for the purposes of linking the time
domain to the spectral regions. This pulse highlights that the current definition of
incremental velocity as the area under the ground acceleration history between zero
crossings is not always appropriate for fully capturing the damage potential of a record.

Figure 5.17: TCU055N record scaled to Sd(T1) = 10 cm: (a) displacement response
spectrum; (b) T@P spectrum; (c) displacement response spectra including spectra for
selected pulse; (d) ground acceleration history.

A comparison of the pulse displacement response spectra shown in Figure 5.17(c)


shows that non-zero (i.e., not at-rest) initial conditions at the onset of the pulse typically
increase the peak displacement response relative to the response obtained using at-rest
Chapter 5: Spectral Shape and Structural Response 132

initial conditions and that the increase can be significant in some cases. As shown in
Figure 5.17(c), the at-rest spectrum of the pulse is relatively flat for T > 3 s. Despite the
shape differences between this pulse and a half-sine pulse, the overall shape of the at-
rest pulse spectrum and the flattening at periods greater than 3td is consistent with the
at-rest pulse spectrum for a half-sine pulse (Figure 5.5). This is illustrated more clearly in
Figure 5.18. Figure 5.18(a) shows the selected ground acceleration pulse and an
equivalent half-sine pulse with the same duration and incremental velocity as the pulse,
and Figure 5.18(b) shows the elastic displacement spectra of the pulses using at-rest
initial conditions. Despite the strikingly different shapes of the pulses in Figure 5.18(a),
the shapes of their displacement response spectra are remarkably similar as shown in
Figure 5.18(b).

Figure 5.18: Pulse from the TCU055N record and an equivalent half-sine pulse:
(a) ground acceleration; (b) displacement response spectrum.

Even when considering the initial conditions imposed by the ground motion, the selected
pulse fails to explain why increasing spectral ordinates are observed in the full motion
between T = 3.2 s and 4 s (Figure 5.17(c)). The T@P spectrum in Figure 5.17(b) shows
that the increasing spectral ordinates in this period range are the result of three different
pulses. (As before, the quotation marks are used to denote that a segment of
acceleration containing positive and negative values is being used.) These pulses start
at t = 26.63 s, 28.78 s, and 23.22 s, respectively. The spectra for two of the three pulses
are shown in Figure 5.19, which demonstrates that these pulses are responsible for the
increasing spectral ordinates between T = 3.2 s and 4 s. For clarity the at-rest pulse
Chapter 5: Spectral Shape and Structural Response 133

spectra are not included in this figure, and only the spectra using the initial conditions
from the ground motion are shown.

Figure 5.19: Displacement spectra of the TCU055N record scaled to Sd(T1) = 10 cm,
including two pulse spectra.

The increasing spectral ordinates for T > T1, a characteristic of the damaging ground
motions in this case study, are created by individual pulses or particular segments of the
acceleration history with significant incremental velocities. These time domain features
are also what cause large inelastic excursions, damage, and possible collapse of the
structure. This particular record is so damaging because it contains multiple pulses with
significant incremental velocities in the same direction (in this case, positive), leading to
a ratcheting-type response in the negative displacement direction. The damaging effects
of these pulses are shown in Figure 5.20, which presents the SDR history of the first story
in Figure 5.20(a) and the unscaled ground acceleration in Figure 5.20(b). The pulses
highlighted in Figure 5.20(b) include the pulses from Figure 5.17 and Figure 5.19. As
Figure 5.20 demonstrates, the significant inelastic excursions occur as the result of the
pulses with significant incremental velocity. Given that these pulses all have positive
incremental velocities, it is not surprising that the structure collapses in the negative
displacement direction.
Chapter 5: Spectral Shape and Structural Response 134

Figure 5.20: TCU055N record: (a) 1st story drift ratio history; (b) ground acceleration
history.

Figure 5.21 shows the 5%-damped displacement response spectra for 60 records used in
the case study, which comprise three groups of records: 20 with the lowest collapse
intensities (Sa(T1)col) (i.e., in the lower tail of the collapse fragility curve) (Figure
5.21(a)), 20 closest to the median collapse intensity (Figure 5.21(b)), and 20 with the
highest collapse intensities (i.e., in the upper tail of the collapse fragility curve) (Figure
5.21(c)). All records are scaled to Sd(T1) = 10 cm. For reference, the mean collapse
intensity of the records in Figure 5.21(a) is Sa(T1)col = 0.55 g compared to mean values of
Sa(T1)col = 1.03 g and 2.37 g for the records in Figure 5.21(b) and Figure 5.21(c),
respectively. These Sa(T1)col values correspond to Sd(T1)col = 24 cm, 45 cm, and 104 cm,
respectively, for Figure 5.21(a, b, and c) and indicate that records in Figure 5.21(c) must
be scaled up over four times, on average, relative to those in Figure 5.21(a) to cause
collapse. In other words, the 20 records in Figure 5.21(a) require an intensity to produce
collapse that is half of the median collapse intensity of all the records in the MRCD137
ground motion set, meaning they can be considered twice as damaging as the 20 records
in Figure 5.21(b). Meanwhile the 20 records in the third group shown in Figure 5.21(c)
need to be scaled to approximately double the intensity of those in the second group
shown in Figure 5.21(b), meaning they are roughly half as damaging as the 20 in the
second group. Similarly, the first group of ground motions (those whose spectra are
Chapter 5: Spectral Shape and Structural Response 135

shown in Figure 5.21(a)) can be considered to be approximately four times more


damaging than those in the third group whose spectra are shown in Figure 5.21(c).

Figure 5.21: Comparison of displacement spectra for records with the (a) lowest;
(b) median; and (c) highest collapse intensities, where all records are scaled to the same
Sd(T1).

A geometric mean spectrum indicated by a thick dashed line is added to each plot in
Figure 5.21 to illustrate that, on average, there is a distinct difference in the spectral shape
of the three groups of records: the lower the collapse intensity, the sharper the increase of
spectral ordinates or, equivalently, the steeper the average slope of the spectra for T > T1.
Does this mean that a record with sharply increasing spectral ordinates for T > T1 will be
damaging (i.e., have a low collapse intensity) to a particular structure? Not necessarily.
As previously discussed, the shape of the spectrum is related to the duration and intensity
of the pulses contained in the ground motion (Figure 5.5 and Figure 5.6) as well as the
state of the system (the initial conditions) when those pulses act on the structure
(Figure 5.8). Sharply increasing spectral ordinates for T > T1 indicate that the record has
one pulse or multiple pulses that impose significant displacement demands on elastic
SDOFs. If the increasing ordinates are primarily due to a single pulse, it is likely the
record will be damaging. If the increasing ordinates are due to a series of pulses,
however, the damage potential will also depend on the signs of the incremental velocities
of these pulses. If the incremental velocities of all or most of these pulses have the same
sign (e.g., all or most are positive), they will tend to push the structure in the same
direction and their damaging effects will be cumulative, leading to a ratcheting-type
behavior. This was the case for the record in Figure 5.20. On the other hand, if some of
Chapter 5: Spectral Shape and Structural Response 136

the controlling pulses have positive incremental velocities while others have negative
values, the structure will likely experience significant inelastic excursions in both
displacement directions, and some pulses might re-center the structure or offset some of
the residual displacement caused by previous pulses. In this case it will typically take a
higher ground motion intensity to collapse the structure compared to a record that
primarily pushes the structure in one direction. Note that, among many other factors, the
effects of cyclic deterioration in the strength and stiffness of structural components can
also influence how a system responds to a pulse as severe deterioration may make the
structure less likely to experience a significant reduction in residual displacement.

An example of a record with significant pulses in opposite directions is the north-south


component of the ground acceleration recorded at the TCU057 station during the 1999
Chi-Chi, Taiwan earthquake, which is denoted TCU057N. This record has the median
collapse intensity (Sa(T1)col = 1.03 g), but it displays sharply increasing spectral ordinates
similar to the damaging records. Figure 5.22 shows that the displacement spectra for this
record and a record with one of the lowest collapse intensities (which appears in Figure
5.21(a)) are very similar. Both records in Figure 5.22 are scaled to Sd(T1) = 10 cm.

Figure 5.22: Comparison of spectra for records with similar spectral shapes but
significantly different collapse intensities.

Figure 5.23, which presents the SDR history of the first story in Figure 5.23(a) and the
unscaled ground acceleration in Figure 5.23(b) for the TCU057N record (the record with
the median collapse intensity in Figure 5.22), shows that there are significant inelastic
Chapter 5: Spectral Shape and Structural Response 137

excursions in both displacement directions. Some pulses offset some of the large inelastic
displacements caused by previous pulses (e.g., the pulse starting just after 45 s that moves
the system in the positive displacement direction after the pulse starting around 44 s
caused significant inelastic displacements in the negative direction). On the other hand,
the record with the low Sa(T1)col in Figure 5.22 experiences significant inelastic
excursions in only one direction and collapses at an intensity of Sa(T1) = 0.59 g. (Note
that the displacement response history to this record is not shown.) This explains why,
despite the similar spectral shapes of the two records shown in Figure 5.22, one record
must be scaled to a much greater intensity than the other to collapse the structure.

Figure 5.23: TCU057N record: (a) 1st story drift ratio history; (b) ground acceleration
history.

Limited investigations were conducted related to predicting which records would be most
damaging from a group of records with similar spectral shapes. These investigations
included comparing the spectrum and T@P due to only positive displacement values to
the spectrum and T@P due to only negative displacement values to provide some insight
into the number of pulses controlling the spectrum and the balance between pulses with
significant positive and negative incremental velocities. The investigations were
inconclusive, and more work is needed in this area.
Chapter 5: Spectral Shape and Structural Response 138

5.6 Measures of spectral shape


Up to this point discussion of spectral shape has focused on the displacement spectrum as
the structural damage and collapse caused by an earthquake are primarily the result of
lateral deformations. Nevertheless, the remainder of this section will focus on the pseudo-
acceleration response spectrum to facilitate direct comparisons between different
measures of spectral shape including epsilon (), which is defined as the number of
logarithmic standard deviations a pseudo-acceleration spectral ordinate of a ground
motion is from the median ordinate predicted by a ground motion prediction equation
(GMPE) at a given period (Baker and Cornell 2005b). Given that a subsequent section
discusses ground motion selection, focusing on the pseudo-acceleration spectrum also
makes sense as many structural engineering design documents use this spectrum and
many earthquake engineering professionals are more familiar with the pseudo-
acceleration spectrum.

To make it clearer how the spectral shape changes from the displacement to the pseudo-
acceleration domain, Figure 5.21 is re-plotted using pseudo-acceleration instead of
displacement. Figure 5.24 presents the 5%-damped pseudo-acceleration response spectra
for 60 records used in the case study: 20 with the lowest collapse intensities (Sa(T1)col)
(Figure 5.24(a)), 20 closest to the median collapse intensity (Figure 5.24(b)), and 20 with
the highest collapse intensities (Figure 5.24(c)). All records are scaled to Sa(T1) = 0.23 g
(Sd(T1) = 10 cm, the same ordinate used in Figure 5.21), and the thick dashed line
indicates the geometric mean spectrum of the records in each plot. Comparison of Figure
5.21 and Figure 5.24 shows that differences in spectral shape are much easier to see in the
displacement domain.
Chapter 5: Spectral Shape and Structural Response 139

Figure 5.24: Comparison of pseudo-acceleration spectra for records with the (a) lowest;
(b) median; and (c) highest collapse intensities, where all records are scaled to the same
Sa(T1).

Despite the fact that, as previously mentioned, the spectral shape is not a perfect predictor
of collapse intensity, a definite trend is observed when comparing the mean spectra in
Figure 5.24 (or Figure 5.21). To quantify the spectral shape a parameter called SaRatio is
used. SaRatio is defined as Sa(T1) normalized by the geometric mean of spectral
acceleration values over a period range. That is

Sa(T1 ) (5.13)
SaRatio =
Saavg (T1 [a, b])

where Saavg(T1[a, b]) is the geometric mean of spectral acceleration values between
periods aT1 and bT1, and a and b are non-negative constants such that a b. Note that
the value of SaRatio does not change when the ground motion is scaled to different
intensities via amplitude scaling. SaRatio is a measure of spectral shape that provides
information about how high or low the value of Sa(T1) is relative to spectral ordinates at
neighboring periods. SaRatio < 1 indicates that the value of Sa(T1) is less than the mean
value of the surrounding spectral ordinates and, depending on the period range used, can
suggest a valley in the spectrum near T1. SaRatio > 1, on the other hand, indicates that the
value of Sa(T1) is greater than the mean value of the surrounding spectral ordinates and,
again depending on the period range used, can possibly indicate at peak in the spectrum
at or near T1.
Chapter 5: Spectral Shape and Structural Response 140

To illustrate the difference in spectral shape that SaRatio quantifies, the spectra of two
records from the 1999 Chi-Chi, Taiwan earthquake are shown in Figure 5.25, where both
records are scaled to Sa(T1) = 0.2 g. The legend in Figure 5.25 indicates the station name
and orientation of each recording. A horizontal line indicating the value of Saavg(T1[0.2,
3]) and a vertical line at T1 are also included in the figure 2. The record in Figure 5.25(a)
has SaRatio = 0.72 and has a spectral valley at T1 while the record in Figure 5.25(b) has
SaRatio = 2.98 but is not in a peak. As one might expect by looking at the difference in
spectral shape, the two records produce very different responses for the case study
structure when scaled to the same value of Sa(T1). The record in Figure 5.25(a) causes the
case study structure to collapse at Sa(T1) = 0.51 g whereas the record in Figure 5.25(b)
does not cause collapse until the intensity reaches more than four times that value at
Sa(T1) = 2.17 g.

Figure 5.25: Pseudo-acceleration spectra for records scaled to a common value of Sa(T1)
with (a) SaRatio < 1; (b) SaRatio > 1.

SaRatio is a direct measure of a records spectral shape because it measures how large
Sa(T1) is relative to ordinates at neighboring periods. Another measure of spectral shape
is , which is an indirect measure or a proxy to spectral shape because it measures the
difference between the Sa(T1) value of a record and the expected (mean) value of Sa(T1)
based on many ground motions with similar causal events (magnitude, fault type, etc.),

2
In this document Saavg is computed using periods uniformly spaced between aT1 and bT1 at intervals of
0.01 s unless noted otherwise.
Chapter 5: Spectral Shape and Structural Response 141

site-to-source distances, and site characteristics. Because compares the actual and
expected values of the spectral ordinate at only a single period, it cannot directly measure
spectral shape.

Figure 5.26 compares the spectral shape of four records as measured by SaRatio and ,
where SaRatio is computed using Saavg(T1[0.2, 3]), and is computed using the 2008
Boore and Atkinson (BA08) GMPE (Boore and Atkinson 2008). The NGA record
sequence numbers are included in Figure 5.26 to identify each ground motion, and the
number in parentheses following the record sequence number indicates which of the two
horizontal components is shown. Additional information about these records is provided
in Table A.2. Figure 5.26(a, b) presents the unscaled spectra for two records with (T1)
-0.8 along with the BA08 predictions. Though these records have similar values their
spectral shapes are very different, which is clearer in Figure 5.26(e) where the records are
scaled to Sa(T1) = 0.2 g. Differences in the width of the peak in the low period range and
the shape of the spectrum for T > T1 are notable, and this difference in spectral shape is
reflected in the difference of SaRatio values: 2.25 vs. 0.89. SaRatios ability to pick up
differences in spectral shape is further illustrated by Figure 5.26(c, d), which shows the
unscaled spectra for two records with (T1) 1.1 but different SaRatios (2.87 vs. 1.21).
As predicted by the difference in SaRatios, the spectral shapes of the records in Figure
5.26(c, d) are very different: one record has high spectral ordinates in the short period
range that generally decrease sharply with an increase in period while the other record
has a relatively flat spectrum for the entire period range shown in Figure 5.26(d). Figure
5.26(e) shows that the records in Figure 5.26(b, d) have similar spectral shapes despite
having very different (T1) values (-0.79 vs. 1.14, respectively). In fact, the spectral shape
of these two records is similar to the mean spectrum of the records with the lowest
collapse intensities shown in Figure 5.24(a) for T > T1. SaRatio captures the similarity in
the spectral shape of the records in Figure 5.26(b, d) as their values of SaRatio are
relatively close (0.89 and 1.21) compared to the values of SaRatio for the records in
Figure 5.26(a, c) (2.25 and 2.87, respectively). The spectral shape of the latter records
(those in Figure 5.26(a, c)) is very similar to the mean spectrum of the records with the
highest collapse intensities shown in Figure 5.24(c) for T > T1.
Chapter 5: Spectral Shape and Structural Response 142

Figure 5.26: Pseudo-acceleration spectra for: (a, b) unscaled negative records; (c, d)
unscaled positive records; (e) records scaled to common value of Sa(T1).

Given that SaRatio was shown to capture differences in spectral shape very well, its
potential to predict collapse intensity is evaluated by computing the correlation between
SaRatio and collapse intensity in the log-log domain (i.e., the correlation between
ln(Sa(T1)col) and ln(SaRatio) is computed). The log-log domain was chosen instead of the
linear-linear domain because the correlation is slightly stronger in the log-log domain and
less sensitive to the period range used to compute SaRatio. Figure 5.27 presents contours
of the correlation coefficient as a function of the a and b values that define the period
range for SaRatio (Equation 5.13). The maximum correlation coefficient between
SaRatio and collapse intensity is = 0.83, which is achieved when periods between
0.2T1 and 3.3T1 are used to compute SaRatio (i.e., when a = 0.2 and b = 3.3). (Note that
this is not the only period range to produce a correlation coefficient of = 0.83.) It is
Chapter 5: Spectral Shape and Structural Response 143

interesting to note that the correlation coefficient is generally much more sensitive to the
upper limit of the period range (the b value) than the lower limit of the period range (the
a value). Including periods up to 3T1 produces good correlation ( 0.80) for all a 1. It
is also interesting to note that the maximum correlation coefficient is achieved when
periods less than T1 are included, which is presumably because the spectrum reflects
pulses in the ground motion and pulses affect the structural response. Though the
spectrum at periods less than T1 may tend reflect pulses that are shorter in duration than
the fundamental period, these pulses can still affect the structural response. This is
particularly relevant for this structure because it is four stories and higher modes
contribute to its response, although it will be demonstrated in Chapter 7 that considering
spectral values at periods less than T1 can also improve response predictions for SDOF
systems (which, by definition, do not have higher modes). Optimum period ranges for
different structures are explored further in Chapter 7.

Figure 5.27: Contour plot of the correlation coefficient () between the natural logarithm
of collapse intensity and the natural logarithm of SaRatio for the case study as a function
of the period range used to compute SaRatio.

It turns out that SaRatio is similar to the Np parameter recently used by Bojorquez and
Iervolino (2011), which they define as the geometric mean of spectral acceleration values
between T1 and some period greater than T1 (they recommend 2T1), normalized by
Sa(T1). SaRatio, which was developed independently of Np, is essentially the inverse of
Chapter 5: Spectral Shape and Structural Response 144

Np except that the period range for computing the geometric mean is not constrained to be
greater than T1.

A scatter plot of the natural logarithm of collapse intensity and the natural logarithm of
SaRatio for each ground motion used in the case study is shown in Figure 5.28, where
Saavg(T1[0.2, 3]) is used to compute SaRatio 3. A linear regression to the data is included
(with the equation shown in Figure 5.28), which shows a reasonable fit with R2 = 0.68.
This plot shows that SaRatio is a very good predictor of collapse intensity. For example,
all ground motions with SaRatio 1 (i.e., ln(SaRatio) 0) are in the lower half of the
collapse fragility curve (i.e., the collapse intensities are less than the median of 1.03 g, or
equivalently less than ln(1.03) = 0.03 as shown in Figure 5.28). However, not all motions
in the lower half of the collapse fragility curve have SaRatio 1 (i.e., ln(SaRatio) 0).

Figure 5.28: Natural log of collapse intensity versus natural log of SaRatio.

For comparison purposes a scatter plot of the natural logarithm of collapse intensity
versus calculated using the BA08 GMPE is shown for all 274 records in Figure 5.29. A
linear fit is also added to the data (with the equation shown in Figure 5.29), which has =
0.32 (R2 = 0.10) and demonstrates that there is much more scatter in the collapse intensity

3
Unless noted otherwise, the remainder of this chapter computes SaRatio using Saavg(T1[0.2, 3]).
Chapter 5: Spectral Shape and Structural Response 145

versus relationship (Figure 5.29) than the collapse intensity versus SaRatio relationship
(Figure 5.28). A review of literature documenting the relationship between (the natural
logarithm of) collapse intensity and shows that the values on the order of 0.4 to 0.6
(R2 from 0.2 to 0.4) have been observed when using record sets without near-fault, pulse-
like motions (Haselton and Deierlein 2008; Liel and Deierlein 2008; Zareian and
Krawinkler 2009; Mousavi et al. 2011). The low correlation between collapse intensity
and observed in this case study may partly be explained due to the presence of near-
fault, pulse-like ground motions in the record set, as it has been noted before that does
not work well for these types of ground motions (Baker and Cornell 2008). However, as
discussed later, even structures subjected to record sets without near-fault, pulse-like
motions can have correlation coefficients smaller than the value observed in this case
study (Figure 5.29).

Figure 5.29: Natural log of collapse intensity versus (T1).

The presence of near-fault, pulse-like ground motions in the record set may also explain
why the trend between collapse intensity and , which is quantified by the slope of the
regression, is not as strong as other studies have found. The slope of the regression
between (the natural logarithm of) collapse intensity and is approximately 0.15 as
shown in Figure 5.29. Haselton et al. (2011) studied 65 reinforced concrete (RC) special
moment-resisting frames (MRFs) using a far-field ground motion set and found that the
average slope was 0.28. They also examined 20 RC ordinary MRFs, which are less
Chapter 5: Spectral Shape and Structural Response 146

ductile than special MRFs, as well as 26 non-ductile RC MRFs and found that the
average slopes were 0.19 and 0.18, respectively, for these buildings, which are values
relatively close to the slope shown in Figure 5.29.

To verify that the correlation with collapse intensity is generally stronger for SaRatio
than for , case studies of over 650 generic moment-resisting frame (MRF), generic shear
wall, and reinforced concrete (RC) MRF structures are used. Details about these
structures, including the structural properties and modeling assumptions, and the ground
motions used are described in Appendix B, but briefly the generic MRF and generic shear
wall structures are those used by Zareian and Krawinkler (2009, Chapter 4) and the RC
MRFs are those used by Haselton and Deierlein (2007, Table 3-3). The generic MRF and
generic shear wall structures range from four to sixteen stories with periods between T1 =
0.2 s and 3.2 s while the RC MRF structures range from one to twenty stories with
periods between T1 = 0.42 s and 2.63 s. Using collapse data generously provided by these
authors, the correlation of collapse intensities with SaRatio and is calculated, and the
results are summarized in Table 5.1. The values are calculated using the BA08 GMPE.
The results demonstrate the correlation with collapse intensity is significantly greater for
SaRatio than for . As seen in Table 5.1 the correlation of the former is on the order of
= 0.8, on average, while the latter is on the order of = 0.5, on average. Less than 3% of
the structures demonstrated a stronger correlation between collapse intensity and than
between collapse intensity and SaRatio. Though not shown, the average slope of the
regression between collapse intensity and is on the order of 0.3 for these structures,
which is consistent with the slope of the RC special MRF structures studied by Haselton
et al. (2011). The average slope of the regression between collapse intensity and SaRatio
is on the order of 1.1 for the structures presented in Table 5.1.

Table 5.1: Summary of correlation between collapse intensity and or SaRatio.


No. of
Data set [ lnSa(T1)col, (1) ] [ lnSa(T1)col, lnSaRatio ]
structures
Mean Range Mean Range
Generic MRFs 396 0.54 0.20 0.75 0.80 0.33 0.95
Generic Shear Walls 252 0.52 0.27 0.65 0.82 0.56 0.96
RC MRFs 30 0.53 0.43 0.64 0.89 0.81 0.93
Chapter 5: Spectral Shape and Structural Response 147

To help understand why some ground motions do not follow the expected relationship
between collapse intensity and (and therefore why and R2 are much lower than those
computed using SaRatio), the pseudo-acceleration spectra are plotted with the BA08
predictions in Figure 5.30 for twelve selected ground motions. The ground motions are
grouped into four types: (1) records with negative and high Sa(T1)col (Figure 5.30(a));
(2) records with neutral and high Sa(T1)col (Figure 5.30(b)); (3) records with neutral
and low Sa(T1)col (Figure 5.30(c)); and (4) records with positive and low Sa(T1)col
(Figure 5.30(d)). Each ground motion in Figure 5.30 has a symbol which corresponds to
the symbol shown in Figure 5.28 and Figure 5.29. The NGA record sequence numbers
are also included in Figure 5.30 to identify each ground motion, and the number in
parentheses following the record sequence number indicates which of the two horizontal
components is shown. Additional information about these records is provided in Table
A.2.

In most cases, the mismatch between the collapse intensity predicted from the regression
with and the actual collapse intensity for the ground motions shown in Figure 5.29 is
the result of failing to capture the spectral shape. A negative value tends to indicate
that the spectrum has a valley at T1 while a positive value tends to indicate that the
spectrum has a peak at T1 (Baker and Cornell 2005b). For records with the same value of
Sa(T1), those with negative values at T1 are expected to have higher spectral ordinates at
other periods and are therefore expected to be more damaging to nonlinear MDOFs
(Baker and Cornell 2005b). However, Figure 5.29 shows that many records do not follow
this trend. Figure 5.30(a) shows that the selected negative records do not have valleys at
T1 and instead have unusually low spectral ordinates with respect to the BA08 prediction
for almost all periods outside of the short period range. Conversely, the positive records
in Figure 5.30(d) do not have spectral peaks at T1 and instead have unusually high
spectral ordinates with respect to the BA08 prediction for almost all periods outside of
the short period range. Some of the neutral records in Figure 5.30(b) and Figure 5.30(c)
actually exhibit peaks and valleys, respectively, at T1 even though a neutral value does
not tend to indicate these features.
Chapter 5: Spectral Shape and Structural Response 148

Figure 5.30: Unscaled pseudo-acceleration spectra with predicted spectral shape from
Boore and Atkinson 2008 (BA08) model for selected ground motions: (a) negative , high
Sa(T1)col; (b) neutral , high Sa(T1)col; (c) neutral , low Sa(T1)col; (d) positive , low
Sa(T1)col.

Because is defined with respect to a GMPE (a prediction of the median) it provides an


indirect measure of median shape, but in some instances (i.e., for some individual
records), such as those shown in Figure 5.30, is unable to infer correct information
Chapter 5: Spectral Shape and Structural Response 149

about the spectral shape. SaRatio, on the other hand, measures the intensity at the
fundamental period relative to intensities at other periods for a given ground motion and
therefore is a direct measure of spectral shape. Comparison of Figure 5.28 and Figure
5.29 shows that for selected records (also shown in Figure 5.30) where was not a good
predictor of Sa(T1)col, SaRatio predicts Sa(T1)col reasonably well.

The performance of an additional spectral shape parameter called eta () is investigated


in Figure 5.31, which shows the natural logarithm of collapse intensity versus . The
parameter was recently proposed by Mousavi et al. (2011) and is a linear combination
of (Sa) and (PGV), where (Sa) is the discussed previously that measures the difference in
spectral acceleration and (PGV) measures the difference between the PGV of a record and
the mean PGV predicted by a GMPE. The parameter is also an indirect measure of
spectral shape. Mousavi et al. studied the relationship between collapse intensity and
for 84 SDOF systems with periods ranging from 0.1 s to 2.0 s and ductility values
between 2 and 12. The SDOFs had a tri-linear backbone curve with zero hardening slope
and a capping ductility equal to 90% of the ultimate ductility. Based on their results, they
formulated as = (Sa) 0.823(PGV) and found that the median value of the correlation
coefficient between and the natural logarithm of collapse intensity was = 0.65 (R2 =
0.42) for these systems. As Figure 5.31 shows, the correlation between and the collapse
intensity obtained for the case study discussed in Chapter 3 is = 0.59 (R2 = 0.35), which
is in good agreement with their findings. It should be noted that the BA08 prediction was
used to compute for the case study whereas Mousavi et al. (2011) used the Campbell
and Bozorgnia (2008) GMPE to compute and calibrate the coefficients of the
equation. Differences in the correlation coefficient due to use of different GMPEs are
expected to be small, however, based on the lack of significant changes to the correlation
coefficient between collapse intensity and when using different GMPEs as well as the
findings of Baker and Jayaram (2008) who observed that correlations between spectral
values at different periods were not sensitive to the choice of GMPE. In summary, the
collapse intensities of the case study showed significantly stronger correlation with than
with ( = 0.59 vs. 0.32, respectively); however, the collapse intensities were most
Chapter 5: Spectral Shape and Structural Response 150

strongly correlated with SaRatio ( = 0.82), which, unlike the other two parameters, is a
direct measure of spectral shape.

Figure 5.31: Natural log of collapse intensity versus .

5.7 Implications for collapse risk assessment


The previous sections illustrated that the spectral shape of a record is related to how
damaging that record is to a particular structure. This finding is consistent with previous
studies such as those discussed in Section 5.2, which showed that the spectral shape of
the records used to analyze a structure can significantly influence the structural response
estimates and the resulting risk estimates (e.g., Baker and Cornell 2006a; Haselton et al.
2011). Given its potential to significantly affect structural response, the spectral shape of
records is an important factor to consider when assessing the collapse risk. To obtain a
good estimate of collapse risk, records with spectral shapes appropriate for the site and
hazard level of interest should be used in response history analyses. The challenge,
however, is determining what the appropriate spectral shapes are.

One method for selecting records with regard to spectral shape is to use a conditional
spectrum (CS) (Lin et al. 2013a; 2013b). A probabilistic seismic hazard analysis (PSHA)
deaggregation, which describes the relative contributions of different magnitude,
distance, and values to the seismic hazard, is used to identify the main causal event that
contributes to the probability of equaling or exceeding a given value of Sa(T). The
Chapter 5: Spectral Shape and Structural Response 151

properties of the main causal or target event (e.g., magnitude, distance, , fault type, etc.)
are used with a GMPE to generate a conditional mean spectrum (CMS), which uses
correlations between values at different periods to compute the expected spectral
values. Generation of the CMS is described in detail by Baker (2011). The CMS, which
describes the expected spectrum given the target event, and the variability of spectral
values at all periods combine to form the CS. The CS can account for multiple causal
events and multiple GMPEs as discussed by Lin et al. (2013a). Additional details about
the CS including information about publicly available software that generates the CS and
selects ground motions matching the distribution of the CS are provided by Lin et al.
(2013b).

The advantage of using the CS to select records is that it describes the full distribution of
the spectrum at all periods (i.e., it captures both the expected spectral shape and the
variability). Parameters such as SaRatio, , or quantify spectral shape by reducing it to a
single value, such as a ratio of spectral ordinates (SaRatio), a normalized difference
between the predicted and observed value of a single spectral ordinate (), or a linear
combination of the former with the normalized difference between the predicted and
observed value of PGV (). Though these parameters can be useful predictors of
structural response, some information about the spectral shape is lost when characterizing
the shape by a single value. Determining the appropriate values of these spectral shape
parameters can also be challenging as standard PSHA deaggregations do not include
information about SaRatio or . Though is a standard output of a PSHA deaggregation,
this parameter is typically significantly less correlated with structural response than either
SaRatio or .

The remainder of this section focuses on the spectral shape parameter SaRatio and
illustrates how it can affect the results of a collapse risk assessment. A method is
proposed for incorporating information about this parameter into collapse risk
assessment. This method is based on modifying the results of response history analyses to
consider SaRatio and is intended for use when ground motions are selected without
Chapter 5: Spectral Shape and Structural Response 152

regard to spectral shape. This method can be useful when there are not enough records to
match the desired spectral shape.

5.7.1 Influence of SaRatio on collapse risk assessment


As discussed in the previous section, the collapse intensity of a ground motion was found
to be strongly correlated with its SaRatio. The SaRatio values of the records used in
response history analysis can have a significant influence on a structures collapse
fragility. As an example consider how the collapse risk assessment results of the case
study structure change when three different record sets are used: (a) the original set of
274 records; (b) half of the original set (137 records) with the lowest values of SaRatio;
and (c) half of the original set with the highest values of SaRatio. Table 5.2 presents the
collapse risk assessment results obtained using these three ground motion sets in terms of
the median and dispersion of the collapse fragility curve, the mean annual frequency of
collapse (c), and the probability of collapse in 50 years (Pc,50), which is the design life of
many typical structures. Because records with lower SaRatios tend to have lower collapse
intensities (i.e., they tend to be more damaging), it is not surprising that the results from
the record set with the lowest SaRatios show that there is a higher risk of collapse relative
to the results from original set that considers all 274 records. What is significant,
however, is the amount by which the collapse risk increases: both c and Pc,50 increase by
approximately 55% (as a percentage of their original values) compared to the original set.
The change in collapse risk is even greater when comparing the set with the highest
SaRatios to the original set: both c and Pc,50 decrease by 65% (as a percentage of their
original values).

Table 5.2: Collapse risk assessment results using different record sets.
Mean Median
Records used SaRatio of Sa(T1)col ln c [10-4] Pc,50*
records [g]
(a) original set (all 247) 1.57 1.03 0.39 3.90 1.93%
(b) half with lowest SaRatio 1.13 0.80 0.26 6.08 2.99%
(c) half with highest SaRatio 2.00 1.33 0.32 1.38 0.69%
*Pc,50 = probability of collapse in 50 years
Chapter 5: Spectral Shape and Structural Response 153

Reductions of 33% and 18% in the dispersion on the median collapse intensity are
observed for the sets with the lowest and highest SaRatios, respectively, compared to the
original set (Table 5.2). This small but noticeable reduction in dispersion is not surprising
because given that SaRatio is a good predictor of collapse intensity, records with greater
similarity in their SaRatios could be expected to have greater similarity in their collapse
intensities, hence less variability and dispersion.

5.7.2 Considering SaRatio in collapse risk assessment


The previous example highlighted the sensitivity of collapse risk assessment results to the
spectral shape of the records as measured by SaRatio. As previously mentioned the
spectral shape of the records should be appropriate for the site and hazard level of
interest, and one method of selecting records is to use the CS, which describes both the
expected spectral shape and the variability. When using records selected to match the CS,
no modification to structural response results is necessary because it is assumed that the
records represent the appropriate distribution of spectral shape. In other words it is
assumed that the structural response results obtained by this method already account for
the effects of spectral shape.

When structural responses are obtained using records selected without regard to spectral
shape, the SaRatios of the records used can significantly impact the collapse fragility
curve and the collapse risk estimate as demonstrated previously. Given that PSHA
deaggregations are not currently available for SaRatio as they are for magnitude,
distance, and , an approximate method for estimating an appropriate or target value of
SaRatio based on available data is proposed. This method is based on computing the
target value of SaRatio from the expected spectral shape based on a PSHA deaggregation.
The structural response results are then adjusted to what is expected when using records
with the target value of SaRatio.

The first step in estimating a target value of SaRatio is obtaining a PSHA deaggregation
for Sa(T1) at the hazard level of interest that shows the relative contribution of different
magnitude, distance, and (M-R-) bins to the seismic hazard at the site. Such
Chapter 5: Spectral Shape and Structural Response 154

deaggregations are available from the USGS (USGS 2012b) or OpenSHA (Field et al.
2003; Field 2013), for example. An expected spectrum is calculated based on the PSHA
deaggregation. As discussed by Lin et al. (2013a; 2013b), the exact CMS, which
considers multiple causal events and GMPEs, can be constructed from the PSHA
deaggregation results, or it can be obtained from the USGS. An approximate CMS can
also be constructed using the mean M-R- values from the PSHA deaggregation.
Comparisons of the exact and approximate CMS for a few example sites are provided by
Lin et al. (2013a; 2013b). The target SaRatio is then estimated from the CMS: that is, the
SaRatio computed from the CMS is taken as the target SaRatio. Note that use of the CMS
is not required as one could compute the target SaRatio from another spectrum as deemed
appropriate.

Figure 5.32 provides an example of how the SaRatio values vary as a function of
magnitude (Mw) and Joyner-Boore distance (Rjb) for different values of (1). Similar to
previous sections, SaRatio is calculated using a period range from 0.2T1 to 3T1, where
T1 = 1.33 s from the case study. The BA08 GMPE is used considering an unspecified
fault type and Vs30 = 285 m/s, which is the average shear wave velocity in the top 30 m of
soil for the case study site. Correlation coefficients between spectral values at different
periods are computed using the equations developed by Baker and Jayaram (2008).

Figure 5.32: SaRatio, computed using a period range from 0.2T1 to 3T1, where
T1 = 1.33 s, as a function of magnitude and distance. SaRatio values are computed from
the CMS based on the BA08 GMPE for an unspecified fault type and Vs30 = 285 m/s.
Different values of (T1) are used to compute the CMS: (a) (T1) = 0; (b) (T1) = 1;
(c) (T1) = 2.
Chapter 5: Spectral Shape and Structural Response 155

When (1) = 0 as illustrated in Figure 5.32(a), SaRatio decreases from approximately 1.7
to 1.3 as the magnitude increases from Mw = 6 to Mw = 8, respectively, and shows a much
smaller variation with distance, particularly for Mw 6.5. Because (1) = 0 the value of
SaRatio shown in Figure 5.32(a) is computed from the median spectrum predicted by the
BA08 GMPE for a given magnitude and distance. Figure 5.32(b) and Figure 5.32(c)
illustrate how the SaRatio values change when (1) = 1 and 2, respectively. Positive
values are chosen as PSHA deaggregations for most sites exhibit positive values at long
return periods of engineering interest (Burks and Baker 2012). The SaRatio values
generally increase as (1) increases because the spectrum becomes more peaked at T1.
For example SaRatio increases from approximately 1.5 to 2.0 for Mw = 7 as (1)
increases from 0 to 2. The regression between collapse intensity and SaRatio in Figure
5.28 implies that the expected value of Sa(T1)col increases by approximately 30% from
1.05 g to 1.36 g as SaRatio increases from 1.5 to 2.0. The values of SaRatio shown in
Figure 5.32 are provided as examples only and should not be generalized to other site
conditions and/or SaRatios computed using a different period range.

As mentioned before, the method previously described for estimating a target SaRatio is
only an approximation. The ideal way to obtain a target SaRatio would be from a
deaggregation that included information about the distribution of SaRatio similar to what
is currently done for where the expected values are reported for each magnitude and
distance bin as well as the overall mean and modal values. This would require
developing a GMPE that incorporates SaRatio and implementing it in standard PSHA
software that provides deaggregations. Given that such GMPEs and deaggregations are
not currently available the method previously described is a reasonable approach for
estimating a target SaRatio from available data. Future research on obtaining appropriate
target SaRatios is necessary.

The following sections describe approximate methods for considering SaRatio in collapse
risk assessment and are based on a knowing the target SaRatio at the hazard level of
interest. The objective of these methods is to obtain a collapse fragility curve that
considers SaRatio and can be used in subsequent collapse risk assessment calculations.
Chapter 5: Spectral Shape and Structural Response 156

The most straightforward method is simply to use ground motions with SaRatios similar
to either the overall target SaRatio or to the target SaRatios from individual events where
the distribution of SaRatios is based on the events relative contribution to the seismic
hazard. This method is simple in that the results of the response history analyses (and
results of the collapse risk assessment based on those analyses) consider SaRatio and do
not need to be modified. A potential downside to this approach comes from simply using
the target value of SaRatio and ignoring the variability in SaRatios. The dispersion in
collapse intensity values may be smaller than what would be found using records with a
greater variety of SaRatios. This was demonstrated, for example, by the results
summarized in Table 5.2 as ground motion sets with a narrower range of SaRatios
produced smaller dispersions in the collapse intensity. The reduced dispersion can have a
significant impact on the resulting collapse risk estimate because, for a given median
collapse intensity, a smaller dispersion will decrease the collapse probabilities in the
lower half of the collapse fragility curve. Because this region of the collapse fragility
curve typically dominates the collapse risk as seen in Chapter 3, the resulting collapse
risk estimate will tend to be smaller (i.e., not conservative) than one found using ground
motions with a wider variety of SaRatios. Another potential downside to this approach is
ensuring the ground motions have SaRatios similar to the target SaRatio, which,
depending on the number of ground motions one desires, may be an additional constraint
on an already very constraining set of ground motion selection criteria.

The other two methods discussed here involve ignoring SaRatio when selecting ground
motions and instead focus the additional effort of considering SaRatio on modifying the
structural response results. These two methods operate under the assumption that the
distribution of SaRatios used to obtain the structural response results provides
information about the results expected at the target SaRatio. The first of these methods is
appropriate when the user wishes to perform a full incremental dynamic analysis (IDA)
(Vamvatsikos and Cornell 2002) in which the collapse intensity of each ground motion is
found by repeating response history analyses with increasing levels of ground motion
intensity until the minimum intensity that causes the structure to collapse is found (the
collapse intensity). The other method is appropriate for a stripe analysis (Jalayer
Chapter 5: Spectral Shape and Structural Response 157

2003) in which ground motions are scaled to a limited number of intensity levels and the
probability of collapse at each intensity is estimated as the fraction of ground motions
causing collapse.

5.7.2.1 Full IDA method


This method is an adaptation of a method proposed by Haselton et al. (2011) to account
for in collapse assessment. As shown in the case study results of Section 5.6, the
collapse intensities are significantly more correlated with SaRatio than with ( = 0.82
vs. = 0.32, respectively), so using SaRatio in place of is expected to provide a better
estimate of the collapse risk. As previously mentioned, this method requires conducting
IDAs to find the collapse intensity of each ground motion in the set, where the ground
motions are selected without regard to SaRatio. A linear regression is performed on the
set of collapse intensity and SaRatio data points in the log-log domain as shown in Figure
5.28 to yield an equation for the expected collapse intensity as a function of SaRatio. The
median collapse intensity considering SaRatio (the modified median collapse intensity)
is obtained from the linear regression using the target SaRatio as the value of SaRatio.

When performing a full IDA, selecting the hazard level at which to obtain the target
SaRatio is not obvious. The relative contribution of each seismic source to the overall
hazard is expected to change with the hazard level, and therefore the target SaRatio is
also expected to change. A reasonable choice would be to select the hazard level
corresponding to the intensity where the c deaggregation reaches its peak because this
hazard level represents the greatest contribution to c. This is the approach Haselton et al.
(2011) suggested for determining the hazard level for selecting the target . This intensity
could be estimated from the c deaggregation computed using the original (unmodified)
collapse fragility curve.

As an example of how to find the modified median collapse intensity, consider the case
study results. The linear regression of the collapse intensities and SaRatios in the log-log
domain shown was performed for the case study in Figure 5.28 and yielded the
relationship ln(Sa(T1)col) = -0.326 + 0.919ln(SaRatio). Suppose a PSHA deaggregation at
Chapter 5: Spectral Shape and Structural Response 158

the intensity with the greatest contribution to c shows that the seismic hazard is
dominated by a magnitude 7 event located 10 km from the site, which is associated with
(T1) = 2.0. The SaRatio of the CMS constructed using these event parameters is equal to
2.0 as shown in Figure 5.32(c). Using the target SaRatio of 2.0 with the regression in
Figure 5.28, the modified median collapse intensity is estimated as ln(Sa(T1)col) = -0.326
+ 0.919ln(2.0) = 0.31, or Sa(T1)col = 1.36 g.

The standard deviation of the median collapse intensity is also affected when SaRatio is
considered. Equation 5.14 gives the reduced conditional standard deviation (ln) of the
modified median collapse intensity (Benjamin and Cornell 1970)

ln = ( res )2 + (m )2 ( SaRatio )2 (5.14)

where res is the standard deviation of residuals from the linear regression between the
collapse intensities and SaRatios in the log-log domain, m is the slope of the regression,
and SaRatio is the standard deviation of the target SaRatio at the hazard level of interest.
The latter quantity presents a challenge as SaRatio is not currently a standard output of
probabilistic seismic hazard deaggregations (whereas standard deviations of magnitude,
distance and are sometimes reported). To overcome this problem it is recommended to
use ln as the standard deviation ln computed from the original (unmodified) collapse
intensities, which is expected to be a conservative approach. Haselton et al. (2011) also
recommend this simplification when using as they observed that the reduction in
dispersion was typically small (e.g., on the order of 10%). In general this magnitude of
reduction in dispersion is not expected to have a significant effect on the collapse risk
estimate. The effect of the reduction in dispersion is expected to be even smaller in this
case as the correlation of collapse intensity with SaRatio is larger and also because the
variability of SaRatio is expected to be smaller. The collapse fragility curve considering
SaRatio is then specified as the lognormal curve with the median equal to the modified
median collapse intensity and the dispersion ln that is computed from the original
(unmodified) collapse intensities.
Chapter 5: Spectral Shape and Structural Response 159

An important factor to consider when determining the collapse fragility curve via a full
IDA is that the same ground motions are used at all intensity levels even though the
events that contribute to the seismic hazard, and thus the ground motion characteristics
(e.g., frequency content and duration), are expected to change with the intensity level. For
this reason and others, including the many approximations involved in this method and
the computational effort required to perform a full IDA, the limited stripe analysis
method described next is preferred over the full IDA method.

5.7.2.2 Limited stripe analysis


In a limited stripe analysis ground motions are scaled to a limited number of intensity
levels and the probability of collapse at each intensity is estimated as the fraction of
ground motions causing collapse. The collapse fragility curve is constructed by fitting a
fragility function (typically lognormal) to these estimated points. The pros and cons of
different fragility fitting methods are discussed by Baker (2013), and the method of
maximum likelihood is shown to be the appropriate method for fitting the results of a
limited stripe analysis. An advantage of limited stripe analysis is that the ground motion
set can change with intensity level to ensure the ground motion characteristics are
consistent with those expected at the intensity level of interest (Jalayer 2003).

The method for considering SaRatio in a limited stripe analysis assumes that the target
SaRatio at the intensity level of interest, Sa(T1) = x, is known. Instead of estimating the
probability of collapse as the fraction of records causing collapse at Sa(T1) = x, the
probability is estimated using a regression on the response history results that describes
the distribution of collapses (and non-collapses) at Sa(T1) = x as a function of
ln(SaRatio). This approach is modeled after a method proposed by Baker and Cornell
(2005b) for analysis using a vector-valued intensity measure and used previously by
Shome and Cornell (1999). The method for considering SaRatio is outlined as follows:
1. Given an intensity Sa(T1) = x, select records without regard to SaRatio and
perform response history analyses with records scaled to Sa(T1) = x.
2. Perform a logistic regression (Agresti 2002) on the collapse state versus
ln(SaRatio) results where the collapse state is 1 if the record caused the structure
Chapter 5: Spectral Shape and Structural Response 160

to collapse and 0 otherwise. The logistic regression gives the probability of


collapse as a function of ln(SaRatio) for Sa(T1) = x. Details about logistic
regression are discussed later.
3. Estimate the probability of collapse at Sa(T1) = x as the probability of collapse
where SaRatio equals the target SaRatio using the logistic regression from step 2.

Logistic regression is commonly used to analyze binary data and is a standard feature of
many statistical software packages. The functional form of the logistic curve used by this
method is

exp(c ln (SaRatio ) + d ) (5.15)


P(C | Sa(T1 ) = x, ln (SaRatio )) =
1 + exp(c ln (SaRatio ) + d )

where C is the collapse state and c and d are coefficients determined by the regression.
Additional information about logistic regression can be found in (Agresti 2002).

The method of considering SaRatio in limited stripe analysis is illustrated by a


hypothetical example using data from the case study. Suppose the target SaRatio is 2.0 at
an intensity of Sa(T1) = 0.7 g, which is an intensity in the peak of the c deaggregation
(i.e., it has a large contribution to c). A logistic regression is fit to the collapse state at an
intensity of Sa(T1) = 0.7 g versus ln(SaRatio) as shown in Figure 5.33. The logistic
regression in Figure 5.33 shows that at the target SaRatio of 2.0 (or, equivalently,
ln(SaRatio) 0.69), the probability of collapse is 0.4%. Therefore, the probability of
collapse at Sa(T1) = 0.7 g considering SaRatio would be estimated as 0.4%. If one had
ignored the effect of SaRatio and estimated the probability of collapse as the fraction of
records causing collapse, a probability of 17.9% would have been found as 49 of the 274
records caused collapse at this intensity. These estimates of the probability of collapse at
Sa(T1) = 0.7 g differ by a factor of more than 40, which can significantly affect the
calculated collapse risk, especially since this intensity was among the most significant
contributors to c.
Chapter 5: Spectral Shape and Structural Response 161

Figure 5.33: Logistic regression on collapse state and ln(SaRatio) for case study data at
Sa(T1) = 0.7 g.

An important caveat to selecting records without regard to SaRatio as stated in step 1 is


that the target SaRatio should be within the range of SaRatios analyzed in order to have a
meaningful regression. Additionally, the analysis results must include collapses and non-
collapses for the regression to be meaningful. To limit the possibility of a few data points
driving the regression, the recommendations of Baker and Cornell (2005b) are adopted: if
two or fewer collapses (or non-collapses) are observed, logistic regression should not be
used and the probability of collapse should be estimated as the fraction of records causing
collapse. The logistic regression may not be meaningful if the SaRatios of collapse and
non-collapse records do not overlap, which is more likely to occur when using a small
number of records. This is particularly significant when the target SaRatio falls between
the SaRatios of the collapse and non-collapse records.

An advantage to estimating the probability of collapse using a regression on SaRatio


rather than as the fraction of records causing collapse is that the estimated probability is
not restricted to be a multiple of 1/N, where N is the number of records analyzed. Because
collapse is a binary state (at least in this context) the fraction of records causing collapse
must be i/N, where i is an integer between 0 and N. If N is large, being restricted to
estimating probabilities of i/N is not a significant issue; however, if N is small this could
Chapter 5: Spectral Shape and Structural Response 162

have important consequences, particularly when the probabilities of collapse are small
and the intensities have significant contributions to c 4. For example, consider a case
where the true probability of collapse at a given intensity is 7.5%, where true means
the probability of collapse computed with an infinitely large number of appropriate
records. When using 20 records, the closest to the true probability one can estimate is
either 5% or 10% (by having either one or two collapses, respectively, out of the 20
records). The contribution to c (the product of the probability of collapse and the slope of
the hazard curve) at this intensity would be, at best, either 1/3 smaller or 1/3 larger,
respectively, than the true contribution. If this intensity has a significant contribution to
c, which is common for intensities in the lower half of the fragility curve, this can
significantly affect the calculated c. This would also affect the contributions of
neighboring intensities whose probabilities of collapse are determined when a collapse
fragility curve is fit to the estimated data points. As the number of points estimated on the
collapse fragility curve is typically substantially much less in a limited stripe analysis
than a full IDA, this could affect a wide range of intensities. When estimating the
probability of collapse based on a logistic regression between the collapse state and
ln(SaRatio), the probability can take any value between zero and one because the logistic
regression is a continuous function.

The method described for considering SaRatio in a limited stripe analysis is applicable
for any type of analysis in which the probability of collapse at a given intensity is
estimated from the results of response history analyses. These include analyses in which
the probability of collapse is estimated at intensities spaced every 0.1 g, for example, or
at intensities corresponding to five different hazard levels. This method is also applicable
when using the approach described in Chapter 4 for estimating the collapse fragility
curve at two intensity levels. This method could even be used with the results of the full
IDA where instead of fitting a fragility curve to the collapse intensities, this method is

4
Note that plotting positions besides i/N can be used to estimate the probability of collapse. For example,
the Blom plotting position (Blom, G. (1958). Statistical estimates and transformed beta variables, John
Willey & Sons, New York.) estimates the probability of collapse as (i-0.375) / (N+0.25). However, the
probability of collapse estimates are still limited to discrete increments (in this case increments of
0.625/(N+0.25) instead of 1/N).
Chapter 5: Spectral Shape and Structural Response 163

used to estimate the probabilities at many discrete intensities (e.g., intensities spaced
every 0.1 g) and the fragility curve is fit to the estimated probabilities.

It is important to remember that both the full IDA and limited stripe analysis methods
discussed here for considering spectral shape in collapse risk assessment require reducing
the spectral shape to a single parameter, in this case SaRatio. As SaRatio is the ratio
between Sa(T1) and the (geometric) mean spectral value over a given period range, it
does not provide information about the relative frequency content of the record at periods
other than T1, so some information about the spectral shape is lost. Furthermore, these
methods focus on adjusting collapse estimates based on a target (expected) value of
SaRatio, which must be estimated as PSHA deaggregations do not provide information
about SaRatio (in terms of either the expected values or the variability). A more direct
way to incorporate spectral shape would be to use records with appropriate spectral
shapes in terms of expected values and variability, such as selecting records to match a
CS based on a PSHA deaggregation. The full IDA and limited stripe analysis methods are
intended for situations when the previously described direct method cannot be used due
to a lack of appropriate records or to adjust results obtained without considering spectral
shape.

5.8 Summary
The relationship between spectral shape and structural response was examined. Because
damage often occurs as a result of large inelastic displacement excursions, the equation of
motion was examined to determine what combinations of structural characteristics and
time domain features of the ground motion lead to large displacement excursions. It was
found that when there is significant area under the ground acceleration history (i.e.,
significant incremental velocity) while a system moves in a direction opposite the ground
acceleration (i.e., the systems velocity and the ground acceleration have opposite signs),
a large displacement excursion is likely to occur, especially if the system yields during
this time. The reason for large displacements occurring when the systems velocity and
ground acceleration have opposite signs is that the momentum the acceleration pulse
imparts to the system is in the opposite direction of the ground acceleration. For example,
Chapter 5: Spectral Shape and Structural Response 164

if the acceleration pulse is positive-valued it imparts momentum in the negative


displacement direction. If the system is moving in the negative displacement direction
(i.e., it has a negative velocity) the momentum from the pulse will add to the movement
of the system and create a larger displacement.

Next, the response to a single half-sine pulse was examined as a ground acceleration
history can be thought of as a series of individual pulses. In the context of this chapter,
the term pulse refers to a segment of the ground acceleration history between zero
crossings. The displacement response spectrum for systems starting from rest subjected to
a half-sine pulse was examined to show that the shape of the spectrum was related to the
time at which individual systems reached their peak displacement (e.g., whether the peak
displacement was reached at the first local maximum/minimum or at subsequent local
maxima/minima). Then the displacement spectrum was generalized for a half-sine pulse
of arbitrary duration and area, and it was noted that it is possible to deduce properties of
the pulse (e.g., the duration td) given only the displacement spectrum. For example, the
spectral ordinates increase sharply for 1 < T/td < 3 (where T is the period of the system)
so from the shape of the displacement spectrum one can infer the duration of pulse that
created that shape. Next the influence of the state of the system (its initial displacement
and velocity conditions) when the pulse starts was examined. It was seen that the initial
conditions can significantly affect the systems response, particularly when td << T, and
that the largest responses typically occur under the initial conditions that maximize the
amount of time the system moves in the direction opposite the ground acceleration (or,
equivalently, in the same direction that the acceleration imparts momentum). The latter
observation is consistent with the findings from examining the equation of motion to
determine what combinations of structural characteristics and time domain features of the
ground motion lead to large displacement excursions.

The peak displacement response to a ground motion was examined next to understand
how the spectral shape arises. It was seen that a smooth spectral region occurs when an
individual segment of the ground acceleration history controls the peak response of a
systems within a given period range. It was also seen that sharp turns or abrupt changes
Chapter 5: Spectral Shape and Structural Response 165

in the shape of the displacement spectrum, such as V-shaped valleys, occur when there
is a change in the segment of acceleration history (i.e. in the critical pulse) that produces
the peak response. Response histories were examined for individual systems in a smooth,
peaked region of the displacement spectrum. Given that the peak response of these
systems was controlled by the same segment of the acceleration history, it was seen that
the reason why one system was at the top of the peak while another system was on the toe
of the peak had to do with the initial conditions of the systems at the start of the
controlling ground motion segment. Differences between elastic and inelastic spectral
shape were also discussed, and it was shown that because prior yielding changes the
inelastic systems initial conditions (relative to the elastic system) at the start of a given
ground motion segment, the peak responses of elastic and inelastic systems can be quite
different from each other. This was illustrated for a system at the top of a peak in the
elastic displacement spectrum where it was seen that the change in initial conditions of
the inelastic system led to a significantly reduced response compared to the elastic
system.

The spectral shape of damaging records (those with low collapse intensities, Sa(T1)col)
was also examined. Results from the case study of Chapter 3 were used to show that the
spectral shape of damaging records is typically distinct from the shape of more benign
records, and that damaging records typically exhibit sharply increasing spectral
displacements for T > T1. Selected damaging records were examined, and it was seen that
individual long duration pulses or segments of the ground acceleration with significant
incremental velocity could form the basis for the spectral shape of the entire record and
even replicate the spectrum of the entire record over large period ranges. This means that
the spectrum can be viewed as a reflection of the duration and intensity of a few critical
individual pulses/segments contained in the ground motion as well as the state of the
systems (the initial conditions) when those pulses/segments act on the various SDOF
systems. This is analogous to the spectrum of the half-sine pulse that was a reflection of
the area and duration of the pulse that created it. The observation about the typical
spectral shape of damaging records was attributed to the fact that these motions contained
pulses/segments with significant intensities and significant durations relative to T1 and
Chapter 5: Spectral Shape and Structural Response 166

that these damaging pulses/segments manifest themselves in the spectral domain with
sharply increasing displacement ordinates for T > T1. Displacement histories of the case
study structure were examined for selected damaging records, and it was seen that the
largest inelastic displacement excursions typically occurred when there were significant
incremental velocities imparting momentum in the same direction the structure was
moving. This is consistent with previous observations about what produces large
displacement excursions based on (a) examining the terms in the equation of motion and
(b) studying how the initial conditions of a system affect its response to a pulse.

Measures of spectral shape were discussed, and three spectral shape measures were
evaluated for their ability to predict the collapse intensity of a record. These measures
included a direct measure of spectral shape called SaRatio, which is defined as the ratio
of Sa(T1) to the geometric mean value of pseudo-acceleration spectral values over a
period range, and two indirect measures of spectral shape, and , that measure the
difference between the spectral ordinate or peak ground velocity of the ground motion
and the mean prediction from a ground motion prediction equations at a given period. It
was seen that for the case study, the collapse intensity showed stronger correlation with
SaRatio than with or . It was theorized that SaRatio performed better because it is a
direct measure of spectral shape rather than an indirect measure of or a proxy to spectral
shape (i.e., SaRatio is computed directly from the spectral values of the records whereas
the other measures are computed with respect to the median of a ground motion
prediction equation).

It was shown that the results of response history analyses can be very sensitive to the
spectral shape of the records used, which is consistent with the findings of other
researchers. To obtain a good estimate of collapse risk, records with spectral shapes
appropriate for the site and hazard level of interest should be used in response history
analyses. Selecting records to match a conditional spectrum (CS) based on PSHA
deaggregations is one approach and is advantageous because it captures both the expected
spectrum and the variability of the spectrum (Lin et al. 2013b). Two approximate
methods for adjusting collapse results obtained using records selected without
Chapter 5: Spectral Shape and Structural Response 167

consideration of spectral shape were presented that are based on using a target value of
SaRatio. One method is appropriate for use with the results from incremental dynamic
analyses in which the collapse intensity of each record is known, and the other method is
appropriate when estimating the probability of collapse conditioned on the ground motion
intensity (where knowing the collapse intensity of each record is not required). Both
methods assume that the distribution of SaRatios used to obtain the analysis results
provides information about the results expected at the target SaRatio. Because seismic
hazard deaggregations do not currently provide information about SaRatio, a method for
estimating a target value of SaRatio was discussed in which the SaRatio is computed
from the spectrum of an event dominating the hazard deaggregation or a mean causal
event. It is stressed that this method is merely an attempt to obtain an estimate and that
more research about obtaining target SaRatios is necessary.
Chapter 6

Intensity Measures for Predicting Collapse

6.1 Overview
This chapter investigates the performance of different ground motion intensity measures
(IMs) in estimating the nonlinear response and particularly the collapse risk of structures.
The performance of various IMs is evaluated using the following criteria: efficiency,
sufficiency, the availability or ease of developing probabilistic seismic hazard
information in terms of the IM, and the robustness of collapse risk estimates produced by
the IM.

A detailed study on the use of pseudo-spectral acceleration averaged over a period range
(Saavg) as a ground motion IM is conducted in this chapter. A case study demonstrates
that Saavg is a better IM for predicting the collapse intensity of a structure compared to
Sa(T1), the pseudo-spectral acceleration at the first-mode period of the structure. Reasons
for this observation are discussed. Note that the 5%-damped spectral ordinates are used
throughout this work unless noted otherwise.

Nearly 700 moment-resisting frame and shear wall structures are used to conduct a
comprehensive evaluation of the sufficiency of Saavg with respect to magnitude, distance,
and spectral shape parameters. Comparisons are made between Saavg and Sa(T1).
Obtaining a seismic hazard curve in terms of Saavg is also discussed.

Finally, the performance of different IMs, specifically Sa(T1), Saavg, and peak ground
velocity (PGV), is evaluated with respect to the collapse risk predictions for a case study.
Particular attention is given to the effects of sufficiency and efficiency on the collapse
risk estimates and the variability of the collapse risk estimates and robustness of the
collapse estimation procedure when different ground motion sets are used to assess the
structural response.

168
Chapter 6: Intensity Measures for Predicting Collapse 169

6.2 Background and literature review


An IM quantifies the intensity of the ground motion and can depend solely on the ground
motion properties (e.g., PGV) or on both the ground motion and structural properties
(e.g., Sa(T1)). Desirable properties of an IM include sufficiency, efficiency, and the
availability of probabilistic seismic hazard information or ease of developing/computing
the seismic hazard associated with the IM. The remainder of this section explains these
properties and provides a literature review of previous studies using IMs to predict
collapse or nonlinear structural response. Particular emphasis is given to IMs related to
spectral values or spectral shape.

As described in Section 2.3, the IM-based approach to collapse risk assessment involves
integrating the collapse fragility curve, which describes the probability of collapse
conditioned on ground motion intensity level, with respect to the seismic hazard curve,
which describes the mean annual frequency of exceeding different ground motion
intensity levels. An implicit assumption in this approach is that structural collapse
depends solely on the intensity level of the ground motion as measured by the IM and not
on other ground motion parameters not captured by the IM. This property is called
sufficiency and means that if, for example, Sa(T1) is used as the IM, the response of the
structure or probability of collapse at a given intensity level is not influenced by
properties such as the magnitude, distance, or duration of the event that produced the
ground motion. Using a sufficient IM is advantageous because it means that ground
motions from a variety of events can be used to assess the response of the structure at a
given intensity level. An issue related to the sufficiency of the IM is whether the
structural response estimates at a given intensity level are the same when using
unscaled/as-recorded ground motions versus ground motions that are scaled to the given
intensity level. The sufficiency of specific IMs is discussed in the literature review.

Another desirable property of an IM is efficiency, which is related to the ability of the IM


to predict the structural response. For example, it is desirable that the peak structural
response exhibits small levels of variability between ground motions with the same
intensity level (as measured by the IM). The efficiency of the IM is also related to the
Chapter 6: Intensity Measures for Predicting Collapse 170

number of ground motions required to estimate the structural response with a given level
of confidence, with more efficient IMs requiring fewer ground motions than less efficient
IMs. For an efficient IM, there is a strong correlation between the ground motion
intensity (as measured by the IM) and the peak structural response and a small variability
of structural response for a given intensity level. The efficiency of different IMs is
discussed in the literature review.

The ability to compute hazard information related to the IM is essential when performing
any type of risk-based structural response assessment using the IM approach.
Specifically, one must be able to develop a ground motion prediction equation (GMPE)
for the IM, that is, one must be able to estimate the probability of exceeding a given
intensity level (as measured by the IM) conditioned on the occurrence of a given seismic
event (e.g., given values of M and R), in order to compute the seismic hazard associated
with the IM. In practice, the availability of and/or the ease of developing probabilistic
seismic hazard analysis (PSHA) tools related to the IM is important (although it should
be noted that some PSHA tools such as OpenSHA (Field et al. 2003; Field 2013) and
CRISIS (Ordaz et al. 2013) allow user-defined GMPEs), as IMs with seismic hazards that
are impractical to compute are unlikely to be very useful, even if the IM has other
desirable properties (e.g., efficiency and sufficiency).

A number of previous studies have examined the ability of given IMs to predict structural
response, including those by Riddell (2007), who studied the correlation of the structural
displacement response, the input energy, and the hysteretic energy of SDOF systems with
23 different IMs, and by Bradley et al. (2010a), who compared the efficiency, sufficiency
and hazard predictability of five different IMs in estimating the drift- and acceleration-
based responses of a 10-story moment-resisting frame (MRF) structure. The following
literature review of previous studies using IMs to predict collapse or nonlinear structural
response is roughly organized according to the type of IM used. Primary focus is given to
IMs related to spectral ordinates and to studies focused on displacement-based response.
It should be noted that many researchers such as Aslani and Miranda (2005a, Chapter 3),
Taghavi and Miranda (2006, Chapter 9), and more recently Bradley et al. (2010a) have
Chapter 6: Intensity Measures for Predicting Collapse 171

noted that IMs that are efficient predictors of displacement-based responses are often not
efficient predictors of acceleration-based responses and vice-versa.

The first IM considered is Sa(T1). Though peak ground acceleration (PGA) was
commonly used to characterize the ground motion intensity many years ago, researchers
such as Shome and Cornell (1999) have demonstrated that Sa(T1) is more efficient than
PGA. Sa(T1) is currently widely used to characterize the response of buildings; however,
Sa(T1) may not be an entirely sufficient IM. For example, Tothong and Cornell (2006)
observed a small dependence between the peak displacement of inelastic single-degree-
of-freedom (SDOF) systems and earthquake magnitude (Mw) when records were scaled to
the same value of Sa(T1); however, many other studies have concluded that Sa(T1) is
sufficient with respect to magnitude and distance (R) for predicting maximum interstory
drift ratios (IDRs) (e.g., Shome et al. 1998; Bradley et al. 2010a). Other studies have
shown that Sa(T1) is not sufficient with respect to a spectral shape parameter called 1
(e.g., Baker and Cornell 2006a; Haselton et al. 2011), meaning that the intensity level at
which the ground motion causes the structure to collapse depends on the value of . In
such cases where the IM is not sufficient with respect to a ground motion parameter,
some modification needs to be made so that the risk assessment results are valid.
Modifications including using a vector IM (e.g., Sa(T1) and ) and careful record
selection are discussed by Baker and Cornell (2005a). Some researchers have also found
that Sa(T1) is not very efficient or sufficient for structures in which higher modes
contribute significantly to the response, such as in tall buildings, and/or for near-fault
ground motions, as the period of ground motion pulses can significantly affect the
response (e.g., Shome and Cornell 1999; Alavi and Krawinkler 2001; Baker and Cornell
2008).

Another issue related to the sufficiency of the IM is whether the structural response
estimates at a given intensity level are the same when using unscaled/as-recorded ground

1
The spectral shape parameter measures the number of standard deviations between the spectral ordinate
at a given period of a particular ground motion and the geometric mean value predicted by a ground motion
prediction equation. More information about this parameter is provided, e.g., by Baker and Cornell (2006a).
Chapter 6: Intensity Measures for Predicting Collapse 172

motions versus ground motions that are scaled to the given intensity level. Luco and
Bazzurro (2007) examined this particular issue with respect to the median drift estimates
obtained for nonlinear SDOF systems and one multiple-degree-of-freedom (MDOF)
system. Their objective was to determine whether, for a given narrow magnitude and
distance (Mw-R) bin and a given level of Sa(T1), using records scaled to the target level of
Sa(T1) produced a bias in the median drift response compared to using unscaled records
with the target magnitude, distance, and unscaled Sa(T1) values. They sorted records into
six Mw-R bins and compared the results obtained from a given target Mw-R bin to the
results obtained using records from a different Mw-R bin that were scaled to match the
Sa(T1) value of the target bin. The authors found that scaled records could introduce
significant bias in the drift response, which they attributed to differences in the median
elastic spectral shape of the different bins. They noted that the bias could be significantly
reduced by selecting records to match the target spectral shape. They also looked at
scaling records within a given bin and found that, on average, records scaled up to a
given intensity level produced larger responses than records that did not need to be scaled
(i.e., that were recorded with the given intensity level).

Baker (2007a) demonstrated that for records scaled to the same Sa(T1) value, the scale
factor did not influence the maximum IDR of a seven-story RC MRF structure (i.e., there
was no bias) when the records were selected based on spectral shape (either via the
parameter or by matching a mean spectral shape). He found that a bias due to scale factor
did exist, however, when records were selected arbitrarily or from a given Mw-R bin.
Bradley et al. (2010a) also observed no bias due to scale factor in the maximum IDR
results of a 10-story RC MRF structure when using records scaled to a common Sa(T1)
value that were selected based on spectral shape.

A number of researchers have been able to improve the efficiency and/or sufficiency of
Sa(T1) by including a secondary parameter that provides information about the spectral
values at periods other than T1. Studies considering the benefit of including information
about spectral values at other periods include those by Shome and Cornell (1999),
Mehanny and Deierlein (2000), and Cordova et al. (2000). Shome and Cornell (1999,
Chapter 6: Intensity Measures for Predicting Collapse 173

Chapter 4) found that including information about the value of Sa(T2), where T2 is the
second-mode period of the structure of interest, for a given value of Sa(T1) decreased the
dispersion of displacement-based responses by approximately 20% for a 20-story
structure. They also found that the relationship between the maximum IDR and the
ground motion intensity was stronger when using a weighted spectral acceleration Sa*
versus Sa(T1), where Sa* is computed as 0.8Sa(T1) + 0.2Sa(T2). Similar results were
found when using an IM that combined the spectral accelerations at the first-, second-,
and third-mode periods of the structure. Mehanny and Deierlein (2000, Chapters 6 and 7)
observed that differences in the ratio of Sa(T2) to Sa(T1) between ground motions could
explain the differences in the structural response. Mehanny and Deierlein also found that
including information about RSa, the ratio of spectral acceleration at an elongated period
to Sa(T1), could improve correlation between the structural response and input ground
motion compared to considering Sa(T1) alone. Cordova et al. (2000) also examined the
use of RSa and, in particular, proposed an intensity measure called S*, which is the product
of Sa(T1) and RSa, where the latter parameter is raised to some power. This power was
recommended to be 0.5, and the elongated period was recommended to be taken as 2T1.
Using these recommendations, S* is equivalent to the geometric mean of the spectral
acceleration ordinates at T1 and 2T1. Using a set of four structures, Cordova et al. found
that S* could significantly reduce the dispersion in maximum IDRs compared to the
dispersion computed using Sa(T1), with reductions of approximately 40% in some cases,
for both ordinary and near-fault ground motion sets.

Vamvatsikos and Cornell (2005b) also studied spectral-based IMs to predict the
maximum IDRs as well as collapse using three structures: a 5-story, a 9-story, and a 20-
story structure that used either MRF or braced frame lateral systems. They first
considered Sa(T), where the period is not necessarily equal to T1, and found that while
some periods could significantly decrease the dispersion in structural response compared
to using T1, the values of these periods were generally difficult to predict a priori. A
notable exception was for structures where the response was dominated by the first mode,
in which case using a period greater than T1 (e.g., 2T1) generally reduced the dispersion
in collapse intensities. Haselton and Baker (2006) reached a similar conclusion while
Chapter 6: Intensity Measures for Predicting Collapse 174

studying the dispersion of collapse intensities of SDOF systems, generally recommending


that a period of 2T1 be used (i.e., using Sa(2T1) as the IM).

Vamvatsikos and Cornell (2005b) also considered using an IM similar to S*, except that
the periods used were not restricted to T1 and an elongated period. Stated somewhat
differently but equivalently in a power-law form, the IM is the product between two
spectral ordinates, where each ordinate is raised to a power. After some optimization,
they essentially recommend that the IM be computed as the geometric mean of spectral
acceleration ordinates at two periods. For the 5-story building, they found that using
periods near T1 and 1.5T1 for the IM resulted in up to a 50% decrease in dispersion of
collapse intensities compared to that computed using Sa(T1). For the 9-story building, the
authors found that the dispersion in collapse intensities is reduced by approximately 25%
using periods near T1 and T2, and for the 20-story building using periods of
approximately 1.4T1 and a period midway between T2 and T3 decreased the dispersion in
collapse intensities by up to approximately 40% compared to the dispersion computed
using Sa(T1). The authors also considered computing the IM as the geometric mean of
spectral acceleration values at three periods and found that while the reduction in
dispersion was marginal compared to that computed using two periods, the resulting
dispersion was much less sensitive to the particular periods used to compute the IM.
Recognizing that including more periods could produce a more robust IM and minimize
the sensitivity of the dispersion to the particular periods used to compute the IM,
Vamvatsikos and Cornell also considered using many spectral ordinates to compute the
IM. Their IM was similar to the geometric mean value of the spectral acceleration over a
period range, but different weights were applied to the spectral ordinates at different
periods. They found that this IM could reduce the dispersion of collapse intensities (as
well as the dispersion of intensities associated with a given level of maximum IDR
response) by up to 50% compared to the dispersion computed using Sa(T1). They found
that the IM performed best when the weights assigned to spectral ordinates increased
with period.
Chapter 6: Intensity Measures for Predicting Collapse 175

A number of researchers have also been able to improve the efficiency and/or sufficiency
of Sa(T1) by including a secondary parameter that indirectly or directly measures the
spectral shape. These researchers include Baker and Cornell (2005b; 2006a), Haselton et
al. (2011), Mousavi et al. (2011), and Bojrquez and Iervolino (2011). Baker and Cornell
(2005b; 2006a) demonstrated that Sa(T1) is not sufficient with respect to the parameter 1,
a proxy to spectral shape that describes the number of logarithmic standard deviations a
spectral ordinate of a given ground motion is from the mean ordinate predicted by a
GMPE at a given period. Note that herein the discussions of refer to the value computed
using the fundamental period of the structure under consideration. Using ordinary (i.e.,
not near-fault, pulse-like) ground motions Baker and Cornell showed that the maximum
IDRs conditioned on a given value of Sa(T1) were correlated with the values of the
ground motions and that risk-based results such as the drift hazard or collapse hazard
obtained using Sa(T1) could be significantly biased unless the records were selected in
accordance with the expected spectral shape or a target value based on a PSHA
deaggregation. Haselton et al. (2011) developed a method for adjusting the collapse
fragility curve when it was computed using records whose values were not consistent
with the target value based on a PSHA deaggregation. Using a case study, they
demonstrated that this method approximated the collapse fragility curve computed using
records whose values were consistent with a PSHA deaggregation. They showed that
not accounting for the effect of could lead to very conservative collapse risk estimates,
as once the effects of were included using their proposed method the median collapse
intensity increased by 60% and the mean annual frequency of collapse decreased by a
factor of 23 for their case study.

Mousavi et al. (2011) used a proxy to spectral shape called , which is a linear
combination of (based on spectral acceleration) and the based on PGV ((PGV)). They
found that is more effective than (based on spectral acceleration) at predicting the
collapse intensity of structures. Mousavi et al. studied the relationship between collapse
intensity and for 84 SDOF systems with periods ranging from 0.1 s to 2.0 s and
ductility values between 2 and 12. The SDOFs had a tri-linear backbone curve with zero
hardening slope and a capping ductility equal to 90% of the ultimate ductility. Based on
Chapter 6: Intensity Measures for Predicting Collapse 176

their results, they computed the parameter as = 0.823(PGV) and found that the
median value of the correlation coefficient between the natural logarithm of collapse
intensity and was approximately = 0.64 for these systems, nearly a 50% increase in
the median correlation coefficient computed using (based on spectral acceleration).

Bojrquez and Iervolino (2011) used a direct spectral shape measure called Np, which is
the ratio of Saavg to Sa(T1), where Saavg is the geometric mean value of spectral
acceleration over a period range. They recommended computing Saavg using periods
between T1 and 2T1. The parameter Np can be viewed as proportional to the average
slope of the spectrum at periods greater than T1. Using case studies of SDOF and MDOF
systems under ordinary, pulse-like, and narrow-band ground motion sets, the authors
found that using a vector IM consisting of Sa(T1) and Np offered significantly improved
predictions of structural response (including maximum IDR, normalized hysteretic
energy demands, and a cumulative damage index) compared to Sa(T1) alone, especially
for the pulse-like ground motions. At intensity levels associated with significant
nonlinearity in the system, this vector IM could reduce the dispersion of maximum IDR
conditioned on intensity by up to 70% compared to using Sa(T1). Bojrquez and Iervolino
also defined a scalar IM called INp, which is the product of Sa(T1) and Np, where Np is
raised to some power, which they recommend as 0.4 based on optimization analyses over
all levels of ground motion intensity. They demonstrated that INp predicted maximum
IDR response significantly better than Sa(T1).

Other researchers including Kennedy et al. (1984), Bianchini (2008), Bianchini et al.
(2009), Kadas et al. (2011), Tsantaki et al. (2012), and Tsantaki and Adam (2013) have
also considered using an average spectral ordinate as an IM. Kennedy et al. (1984)
studied the response of SDOF systems and noted that the level of inelastic structural
response could be better predicted using an average spectral value over periods larger
than the fundamental period of the structure compared to using the spectral ordinate at a
single, effective frequency. They noted that the median predictions were only slightly
better when using the average ordinate; however, they did find a reduction in the
variability of the results, which they attributed to the smoothing effect of averaging
Chapter 6: Intensity Measures for Predicting Collapse 177

compared to using a single ordinate that could be sensitive to local peaks and valleys in
the spectrum. Kadas et al. (2011) considered the integral of spectral acceleration over a
period range between an initial effective period and a lengthened period as an IM to be
used for ground motion selection. For a set of seven short- and medium-period frame
structures, they found that the ground motion intensity measured by this IM was well-
correlated with the maximum IDRs (a range of IDRs from approximately 0.01 to 0.02
radians was considered), with an average correlation coefficient of = 0.85 for the
structures considered. Bianchini (2008) and Bianchini et al. (2009) studied the
displacement response of SDOF and MDOF systems with ductility levels () ranging
from = 1 to 6 using Saavg as the IM. The authors found that the minimum dispersion in
the results was produced using an upper limit on the period range of approximately 2T1
for low ductility levels and approximately 3T1 for high ductility levels ( 4). For the
lower limit of the period range, they recommended using T1 for SDOF systems and
approximately 0.25T1 for MDOF systems with significant contributions from higher
modes. Tsantaki et al. (2012) studied the collapse of non-degrading bilinear SDOF
systems vulnerable to P- effects. The periods of the SDOFs ranged from T1 = 0.1 s to
5.0 s, and various hysteretic behaviors and effective post-yield stiffness values were
considered. They found that Saavg computed between T1 and an elongated period reduced
the dispersion in collapse intensities between 25% and 45% compared to the dispersion
found using Sa(T1). The recommended upper limit of the period range was between
1.2T1 and 2.1T1 and varied as a function of the period of SDOF. The authors found that
Saavg did not reduce the dispersion compared to Sa(T1) for systems with T1 = 0.1 s. In a
related study using the same SDOFs but focusing on systems with bilinear hysteretic
behavior, Tsantaki and Adam (2013) recommended using an upper limit of 1.6T1 for
systems with T1 0.15 s. They found that the optimum upper limit of the period range
was not very sensitive to the damping ratio or the effective post-elastic stiffness.

The inelastic spectral displacement at the fundamental period of the structure (Sdi(T1)),
has also been considered as an IM. Aslani and Miranda (2005a, Chapter 3) examined the
efficiency of four different IMs, including Sd(T1), Sdi(T1), PGA normalized by 42/T12,
and S*. It should be noted that Sd(T1) is the elastic spectral displacement and is directly
Chapter 6: Intensity Measures for Predicting Collapse 178

proportional to Sa(T1), such that the dispersion of structural responses measured by either
IM is identical. The inelastic spectral displacement was calculated using an elastic-
perfectly-plastic SDOF system with a period equal to the fundamental period of the
structure under consideration. The yield displacement of the SDOF was determined based
on a static pushover analysis of the MDOF structure. They found that Sdi(T1) was the
most efficient IM for predicting both the maximum IDR at different stories of a seven-
story building and the overall maximum IDR in the structure for a wide range of intensity
levels, although at higher levels of intensity the performance of S* was comparable.
Ruiz-Garca and Miranda (2005, Chapter 7) also found that Sdi(T1) was more efficient
than Sd(T1) when computing the maximum and residual IDRs, even for long-period
structures.

Luco and Cornell (2007) investigated a series of IMs based on Sd(T1) and/or Sdi(T1),
some of which also incorporated Sd(T2), for predicting the maximum IDR. Note that 5%
damping was not necessarily used in calculating the spectral ordinates in their study.
They considered steel MRF structures of 3, 9, and 20 stories and both ordinary and near-
fault ground motion sets. They found that IM1I&2E, which incorporates Sd(T1), Sdi(T1),
Sd(T2), and the modal participation factors of the first and second modes of the structure,
was generally the most efficient and sufficient IM for all structures and ground motion
sets considered. An exception to the previous statement is for the 20-story structure under
the near-fault ground motions, for which a dependence on distance was observed.
Tothong and Luco (2007) also considered Sdi(T1) and found that for first-mode-
dominated structures, the drift-hazard and probability of collapse conditioned on this IM
were generally not dependent on the value of the ground motion used (opposed to
Sa(T1), which showed a dependence on ). They also used IM1I&2E for structures that were
not first-mode dominated and reached similar conclusions about the efficiency and
sufficiency of this IM.

Akkar and Miranda (2004) found that inelastic displacement ratios (the ratio of peak
inelastic to elastic displacement) of short-period SDOFs were sensitive to the ratio of
spectral velocity at the fundamental period to the PGV, particularly for weak structures
Chapter 6: Intensity Measures for Predicting Collapse 179

with lateral strength ratios greater than four. By incorporating this spectral ratio into the
prediction of inelastic displacement ratios (and therefore peak deformations of short-
period inelastic systems), they were generally able to reduce dispersion in the results by
at least 50%.

Researchers have also considered using PGV as an IM. Akkar and zen (2005) studied
deformation demands in SDOF systems with periods between 0.1 s and 4.0 s subjected to
near-fault records that did not exhibit strong pulse effects. They found that for moderate
to large levels of nonlinearity, PGV was typically significantly more correlated with the
deformation demand than PGA, PGV/PGA, and Sa(T1), especially for short-period
structures. They found that the dispersion of inelastic displacement ratios measured by
PGV and Sa(T1) were generally comparable, except for short-period structures with large
strength reduction factors, for which PGV yielded lower dispersions.

Bradley et al. (2010a; 2010b) studied PGA, PGV, Sd(T1), Sdi(T1), and SI as IMs for
assessing the structural response of a 10-story RC MRF structure. SI is the spectrum
intensity (Housner 1952) and is computed as the integral of pseudo-spectral velocity
between periods of 0.1 s and 2.5 s. They found that while PGV and SI produced
marginally higher dispersions on the collapse intensities compared to Sd(T1) and Sdi(T1),
the former IMs produced annual probability of collapse estimates that were significantly
smaller than those produced by Sd(T1) and Sdi(T1). The authors attributed this to the fact
that PGV and SI were significantly easier to predict from a GMPE perspective (i.e., they
had less dispersion in the expected intensity conditioned on magnitude and distance).
Similar conclusions were reached when considering the drift and acceleration hazards.

Kurama and Farrow (2003) examined the ability of several IMs to reduce scatter in the
peak lateral displacement demands of SDOF and MDOF systems. After scaling ground
motions to the mean IM value of the ground motion set, they computed the peak inelastic
displacement demands and compared each IM using the coefficient of variation (COV) of
the demands, which is the standard deviation normalized by the mean. They found that
spectral-based IMs such as Sa(T1) and Saavg (computed between T1 and a lengthened
Chapter 6: Intensity Measures for Predicting Collapse 180

period) generally resulted in the smallest COVs for SDOF systems. However, they noted
that the maximum incremental velocity (MIV), which is the area under the ground
acceleration between two consecutive zero crossings, generally produced the lowest
COVs for SDOF systems with periods between approximately 0.1 s and 0.5 s and large
strength reduction factors (R 6) and for systems on soft soil sites with fundamental
periods between approximately 0.2 s and 1.4 s and R 4. The authors also found that
MIV produced smaller COVs in the peak roof displacement and IDRs compared to Saavg
for four-story and an eight-story RC MRF structures using records obtained from soft soil
sites. A GMPE for MIV has also been developed (Guamn 2010).

6.3 Spectral acceleration averaged over a period range (Saavg) as an IM


This section explores using the spectral acceleration averaged over a period range (Saavg)
as an IM for collapse risk assessment. Saavg is computed as the geometric mean of
spectral acceleration values between periods T1 and TN

1/ N
N
Saavg (T1, , TN ) = Sa (Ti ) (6.1)
i =1

where N is the number of periods used to compute Saavg. Note that in Equation 6.1 the
subscripts on the periods do not indicate the modal periods (e.g., Ti for i = 1 is not
necessarily equal to the first mode period). A period range between 0.2T1 and 3T1 with
a uniform period spacing of 0.01 s will be used to compute Saavg throughout this chapter
unless noted otherwise. The effect of period range and spacing is studied in Chapter 7.

Note that while the geometric mean and arithmetic mean values are generally very
similar, the former has some advantages for computing the average spectral acceleration
as it is less sensitive to extreme spectral ordinates (i.e., very high or very low), which
may significantly influence the arithmetic mean value. More importantly, GMPEs for
Saavg that are used in probabilistic seismic hazard analysis and the computation of seismic
hazard curves can be computed based on existing GMPEs for spectral acceleration at a
Chapter 6: Intensity Measures for Predicting Collapse 181

single period when Saavg is computed using the geometric mean. Further discussion on
obtaining GMPEs and hazard curves for Saavg is provided in Section 6.3.4.

6.3.1 Relationship between Saavg,col, Sa(T1)col and SaRatio


SaRatio, as defined in Equation 5.13, is the ratio of Sa(T1) to Saavg. As shown in Section
5.6, a strong linear relationship exists between Sa(T1)col and SaRatio in the log-log
domain. For the purposes of this section the relationship is examined in the linear-linear
domain and is of the form

Sa(T1)col = mSaRatio (6.2)

where m represents the slope of the line. In the linear regression analysis the ordinate
Sa(T1)col is constrained to be equal to zero when SaRatio = 0 (i.e., the fitted straight line
passes through zero). An example of the linear trend between Sa(T1)col and SaRatio is
presented in Figure 6.1, based on data from the case study described in Section 3.5. This
figure also includes the line and equation of the form shown in Equation 6.2 found
through linear regression of the data.

Figure 6.1: Sa(T1)col versus SaRatio with linear regression through the origin.

As shown in Figure 6.1, the data points do not lie perfectly on the regression line and are
scattered around it. The degree of scatter around this line is related to the efficiency of
Chapter 6: Intensity Measures for Predicting Collapse 182

Saavg as predictor of collapse intensity. This can be understood by examining Equation


6.2. Rearranging Equation 6.2 in terms of m yields

Sa(T1 )col
m= (6.3)
SaRatio

Rewriting Sa(T1)col in terms of the unscaled (i.e., as-recorded) value of Sa(T1) and using
the definition of SaRatio yields

SFcol Sa(T1 )unscaled


m= (6.4)
Sa(T1 )unscaled

Sa avg ,unscaled

where SFcol denotes the minimum factor by which the (unscaled or as-recorded) ground
motion must be scaled to trigger the structure to collapse. In other words, SFcol is the
collapse scale factor and denotes the ratio between Sa(T1)col and Sa(T1)unscaled. Equation
6.4 simplifies to

m = SFcol Sa avg ,unscaled = Sa avg ,col (6.5)

Equation 6.5 indicates that if all data points in a plot of Sa(T1)col versus SaRatio were to
lie exactly on the regression line (i.e., if there were no scatter) Saavg would be a perfect
predictor of collapse intensity. This means that all ground motions would cause the
structure to collapse when the ground motion intensity meets or exceeds a certain value
of Saavg, which per Equation 6.5 equals the slope of the regression, m. In reality Saavg is
not a perfect predictor of collapse intensity, which is reflected by the scatter about the
regression line shown in Figure 6.1. The scatter around the straight line shown in Figure
6.1 then indicates that the ratio between Sa(T1)col and SaRatio changes from one ground
motion to another (i.e., exhibits record-to-record variability). The better or more efficient
Saavg is at predicting the collapse intensity, the less variability in the values of Saavg,col and
the closer the data points are to the regression of Sa(T1)col versus SaRatio.
Chapter 6: Intensity Measures for Predicting Collapse 183

In general, the median value of Saavg,col from the collapse fragility curve and m, the slope
of the regression between Sa(T1)col and SaRatio, are expected to be similar because both
measures describe the best fit to the collapse intensity data; however, they are not
necessarily expected to be exactly equal to each other because they have different
definitions of the best fit. The median value of Saavg,col is determined by the method of
maximum likelihood, which seeks the lognormal distribution parameters (i.e., the median
collapse intensity and the dispersion values) that maximize the likelihood of observing
the collapse intensity values given the candidate values of the lognormal distribution
parameters. The slope of the regression between Sa(T1)col and SaRatio, on the other hand,
is determined using ordinary linear least squares regression, which seeks the slope that
minimizes the sum of the squared residuals (i.e., the differences between the observed
values of the data points and the values predicted by the regression). A comparison of
these values for the case study discussed in Section 3.5 illustrates their similarity: the
median value of Saavg,col is 0.70 g as shown in Figure 6.2(b) while the slope of the
regression between Sa(T1)col and SaRatio is 0.71 g as shown in Figure 6.1.

Figure 6.2: Collapse fragility curve for (a) IM = Sa(T1); (b) IM = Saavg. The circles are the
collapse intensities of individual ground motions, and the solid lines are the lognormal
fragility curves fitted to the data using the method of moments.
Chapter 6: Intensity Measures for Predicting Collapse 184

6.3.2 Why Saavg is more efficient than Sa(T1) in predicting the collapse
intensity
Figure 6.2 compares the collapse fragility curves using two different IMs for the case
study discussed in Section 3.5: Figure 6.2(a) uses Sa(T1) while Figure 6.2(b) uses Saavg.
As shown in this figure, the use of Saavg as an IM results in a significant decrease in the
dispersion (ln or ) of the collapse intensities compared to the value computed when
Sa(T1) is used as IM: lnSaAvg = 0.22 versus lnSa(T1) = 0.39, which corresponds to a 44%
reduction in dispersion.

Though the median collapse intensities are different for the two IMs shown in Figure 6.2
(0.70 g for Saavg versus 1.03 g for Sa(T1)), the smaller dispersion when using Saavg is
unaffected by this difference due to the lognormal distribution of the data. For example, if
a scale factor of 1.03/0.70 were applied to all Saavg,col values to make the median collapse
intensity the same for both IMs, the dispersion of the scaled Saavg,col values would still be
lnSaAvg = 0.22.

To illustrate the large reduction in the dispersion of collapse intensities when Saavg is
used as the IM, Figure 6.3 presents a scatter diagram of collapse intensities for each
individual record when using each of the two IMs. Figure 6.3(a) shows the collapse
intensities when using Sa(T1) as IM, while Figure 6.3(b) shows the collapse intensities
when using Saavg. The significant reduction in scatter around the median value (the
horizontal black line) is apparent when using Saavg.
Chapter 6: Intensity Measures for Predicting Collapse 185

Figure 6.3: Collapse intensity scatter plots for (a) IM = Sa(T1); (b) IM = Saavg. The
horizontal black line indicates the median value.

The colors of the dots in Figure 6.3 correspond to the value of SaRatio for each ground
motion. SaRatio, as previously defined in Chapter 5, is the ratio of Sa(T1) to Saavg. Figure
6.3(a) shows that there is strong relationship between Sa(T1)col and SaRatio, with small
values of SaRatio (i.e., dark blue dots) generally indicating small values of Sa(T1)col
(damaging records that trigger collapse at low levels of Sa(T1)) and large values of
SaRatio (i.e., orange and red dots) generally indicating large values of Sa(T1)col (more
benign records that will not produce collapse unless they are scaled to have very large
values of Sa(T1)). This indicates that the spectral shape of a record, as captured/measured
by SaRatio, affects the collapse intensity of a record when Sa(T1) is used as an IM. As
discussed in Sections 5.6 and 5.7 this has important implications when computing the
collapse fragility curve and the collapse risk because the SaRatios of the ground motions
used to compute the structural response can bias the results. Methods for incorporating
the effect of SaRatio when using Sa(T1) as an IM were discussed in Section 5.7. Figure
6.3(b), meanwhile, indicates that there is no discernible relationship between SaRatio and
Saavg,col. This suggests that Saavg is sufficient with respect to SaRatio (this will be
examined more formally in Section 6.3.3.3) and accounts for spectral shape features
affecting structural response that Sa(T1) misses.

The previous case study is not the only evidence that Saavg is a more efficient predictor of
collapse intensity than Sa(T1). For example, Tsantaki et al. (2012) studied bilinear non-
deteriorating SDOF systems with periods ranging from T1 = 0.2 s to 5.0 s that were
Chapter 6: Intensity Measures for Predicting Collapse 186

vulnerable to the P- effect. Using a period range of T1 to T1+T1 to compute Saavg


(where T1 varied between 0.2 s and 1.2 s, with larger values of T1 associated with
larger values of T1), Tsantaki et al. found that, on average, Saavg reduced the dispersion in
collapse intensities by at least 25% and up to 45% compared to Sa(T1).

Over 650 generic moment-resisting frame (MRF), generic shear wall, and reinforced
concrete (RC) MRF structures were used as part of this study. Details about these
structures including the structural properties and modeling assumptions and the ground
motions used are described in Appendix B, but briefly the generic MRF and generic shear
wall structures are those used by Zareian and Krawinkler (2009, Chapter 4) and the RC
MRFs are those used by Haselton and Deierlein (2007, Table 3-3). The generic MRF and
generic shear wall structures range from four to sixteen stories with fundamental periods
between T1 = 0.2 s and 3.2 s while the RC MRF structures ranged from one to twenty
stories with fundamental periods between T1 = 0.42 s and 2.63 s. Using collapse data
generously provided by these authors, the collapse intensities reported in terms of
Sa(T1)col were converted to Saavg,col by back-calculating the scale factor required to
produce collapse for a given record and then multiplying the Saavg value of the unscaled
record by this factor. Collapse fragility curves were generated for each structure and the
dispersion on the collapse intensities was recorded.

A summary of results from these case studies provided in Table 6.1 shows that using
Saavg reduces the dispersion in collapse intensities by an average of approximately 40%
for generic MRF and shear wall structures and over 50% for RC MRF structures
compared to dispersions computed when using Sa(T1). In addition to being more efficient
on average, Saavg was also more efficient than Sa(T1) for all but four of the structures
examined. For these four structures lnSaAvg was no more than 6% greater than lnSa(T1).
These results provide compelling reasons for using Saavg as an IM instead of Sa(T1) if
efficiency of the IM when estimating structural collapse is a primary concern. Discussion
about the effects of structural system, fundamental period, and period range used for
averaging on the efficiency of Saavg are provided in Chapter 7.
Chapter 6: Intensity Measures for Predicting Collapse 187

Table 6.1: Summary of dispersion on the collapse fragility curve for Saavg and Sa(T1).
No. of ln ln Reduction in ln
Data set
structures Saavg Sa(T1) using Saavg (%)*
Mean Mean Mean Range
Generic MRFs 396 0.22 0.38 40 (-6) - 66
Generic Shear Walls 252 0.28 0.48 41 17 - 67
RC MRFs 30 0.20 0.44 54 40 - 63
* Difference between lnSa(T1) and lnSaAvg, expressed as a percentage of lnSa(T1)

So why is Saavg more efficient at predicting the collapse intensity than Sa(T1) (i.e., why is
there less dispersion in the collapse intensities when using Saavg as an IM)? Sa(T1)
measures the pseudo-acceleration of an elastic SDOF oscillator. As such, it is, by
definition, the best predictor for estimating the peak displacement (or pseudo-
acceleration) response of a 5%-damped linear elastic oscillator and therefore can be a
good predictor of deformation response parameters for MDOF structures that are
dominated by the first mode and have linear behavior or exhibit low levels of
nonlinearity. However, Sa(T1) does not provide information about the peak (elastic)
response of oscillators with other periods that might provide information and may be
important when estimating the structural response of strongly nonlinear structures.
Meanwhile, Saavg contains information about peak response of oscillators at other
periods, and therefore it provides additional information that is valuable when estimating
the collapse of structure. For example, periods between 0.2T1 and 3T1 were used to
compute Saavg for the case study, and this period range includes the first and second
modes of the structure. Sa(T1) also does not consider how the response changes once the
system yields (i.e., the inelastic response). When a system yields the initial conditions
(i.e., the deformation and velocity) at the onset of severe long-duration ground motion
pulses change relative to those in the elastic system, and the response can be very
different from the elastic response. This is especially true if the elastic ordinate at the
fundamental period is at or very near a spectral peak or a spectral valley, as the initial
conditions of the oscillator when hit by major ground motion pulses are a major factor in
the creation of these spectral features. As discussed in Chapter 5, the response of a
system to a pulse can be very sensitive to the state of the system (i.e., the displacement
and velocity) at the initiation of severe pulses. While Saavg is also based on elastic
Chapter 6: Intensity Measures for Predicting Collapse 188

ordinates, the influence of initial conditions is minimized by averaging the response over
a range of periods, and therefore also encompassing a variety of initial conditions.

As demonstrated in Chapter 5, the spectral shape of the ground motion record is strongly
correlated with the collapse intensity when using Sa(T1). When the records were scaled to
a common value of Sa(T1), records with sharply increasing spectral ordinates for T > T1
were more damaging (i.e., they had lower Sa(T1)col values) than records with relatively
constant or decreasing spectral ordinates for T > T1. The increasing spectral shape was
shown to be the result of an acceleration pulse or series of pulses in the record with
significant incremental velocities combined with the initial conditions of the elastic
systems at the onset of the pulse(s). These pulses were also shown to produce inelastic
displacement excursions in the structure which can lead to ratcheting (accumulating
displacements in primarily one direction) and cause sidesway collapse. By taking an
average spectral value over a range of periods, Saavg accounts directly for spectral shape
features and indirectly for the underlying pulses in the ground motion that cause damage.
Stated somewhat differently, the response spectrum can be viewed as a signature of the
duration and intensity of a few individual ground motion pulses or pulse segments within
the ground acceleration combined with the initial conditions/state of the system when
the pulse hits the oscillator at its base. Saavg incorporates more information about the
ground motion by taking the average value of the signature over several points as
opposed to Sa(T1), which takes the value of the signature at a single period and
therefore is strongly affected by the state of the system at the onset of the pulse. The
additional information about the record that Saavg incorporates is what makes it a more
efficient measure of collapse intensity than Sa(T1).

6.3.3 Evaluating the sufficiency of Saavg


This section evaluates the sufficiency of collapse intensities measured by Saavg with
respect to magnitude, distance, and spectral shape. Essentially, this section explores
whether some ground motion parameter influences the value of Saavg,col (e.g., whether
records from large magnitude events are likely to produce either larger of smaller values
of Saavg,col compared to records from smaller magnitude events). It is desirable that the
Chapter 6: Intensity Measures for Predicting Collapse 189

values of Saavg,col be conditionally independent of (i.e., not affected by) the values of the
ground motion parameters that are already incorporated in estimating the intensity
measure in the seismic hazard analysis so that no bias exists in the structural response
results due to the ground motions selected for response analysis when scaled to common
values of the intensity measure.

The sufficiency is evaluated by performing a standard linear regression of the ln(Saavg,col)


values with the ground motion parameter of interest to identify possible trends (e.g.,
systematic increments or reductions) in the response parameters computed using records
with different ground motion parameters. The regression is of the form

E[Y | X = x] = 0 + 1x (6.6)

where Y is taken as ln(Saavg,col), X is the ground motion parameter of interest, x is some


value of X, 0 and 1 are coefficients determined by the regression, and E[Y | X = x] is the
expected value of Y given some value x. A hypothesis test is then performed with the null
hypothesis being that 1 = 0 (i.e., that the expected value of Y does not depend on the
value of X). The probability of observing a 1 value at least as large (in absolute value
terms) as the 1 value found by the regression given that the true value of 1 equals 0 is
calculated. This probability is called the p-value and is used to determine whether the
null hypothesis can be rejected at some predefined significance level. A 5% significance
level is typically used when judging the sufficiency of an IM (e.g., Baker and Cornell
2005b; Luco and Cornell 2007; Bradley et al. 2010a), meaning that a p-value of less than
0.05 indicates that the IM is not sufficient with respect to the ground motion parameter of
interest. The 5% significance level is also used here. The natural logarithm of the collapse
intensities is used because the linear regression and computation of the p-value assume
that Y is normally distributed (e.g., Benjamin and Cornell 1970). The collapse fragility
curve seen in Figure 6.2(b) shows that the Saavg,col values are lognormally distributed,
meaning that ln(Saavg,col) is normally distributed. The lognormal distribution of the
Saavg,col values is also confirmed by the Kolmogorov-Smirnov goodness-of-fit test (see,
e.g., Benjamin and Cornell 1970) at the 5% significance level.
Chapter 6: Intensity Measures for Predicting Collapse 190

6.3.3.1 Magnitude
Figure 6.4 shows the natural logarithm of the collapse intensities measured by Saavg
plotted with respect to the earthquake magnitude (Mw). An ordinary linear least-squares
regression reveals a slope of -0.07 with a p-value of 0.14, indicating that collapse
intensities are sufficient with respect to magnitude. Note that the range of magnitudes
considered is rather small (Mw from 6.93 to 7.62) and a more thorough evaluation would
consider a wider range of magnitudes. In ordinary linear least-squares regression, each
data point is weighted equally. Because 168 of the 274 ground motions are from the Mw =
7.62 Chi-Chi earthquake, the ordinary regression is heavily influenced by how well it fits
the Chi-Chi data. To determine how significantly this affected the regression, weighted
regressions were performed and the sufficiency was reevaluated. The data points were
weighted so that specified magnitude windows have equal influence on the regression.
The weighted regression was repeated using different magnitude windows to test the
sensitivity of the results, but none of the windows tested produced p-values smaller 0.05,
which suggests that the values of Saavg,col are sufficient with respect to magnitude.

Figure 6.4: ln(Saavg,col) plotted against earthquake magnitude with the ordinary linear least
squares regression of the data.

The fact that the p-values of the regressions are somewhat close to 0.05 (the threshold
between sufficiency and insufficiency) is not surprising given that earthquake magnitude
Chapter 6: Intensity Measures for Predicting Collapse 191

influences the frequency content and thus the spectral shape of a ground motion. This is
reflected in ground motion prediction equations (GMPEs), which use a magnitude-
dependent term to predict the response spectrum. Figure 6.5 shows the response spectra
predicted by the 2008 Boore and Atkinson (Boore and Atkinson 2008) GMPE, denoted
BA08. Spectra are computed for Mw = 6.5, 7.0, and 7.5 conditioned on a Joyner-Boore
distance (Rjb) of 12 km, which represents the average Rjb of ground motions used in the
case study, a shear wave velocity of Vs30 = 285 m/s, which represents the shear wave
velocity at the case study site, and an unspecified fault type. The unscaled spectra in
Figure 6.5(a) show that the spectral ordinates at a given period increase with magnitude,
as expected, and that the width of the spectral peak in the short period range also
increases with magnitude. Figure 6.5(b) shows the same spectra scaled to a common
value of Saavg (equal to 0.12 g), where the dashed vertical lines bracket the period range
used to compute Saavg (from 0.2T1 and 3T1, where T1 = 1.33 s from the four-story steel
case study building). Saavg is the geometric mean of spectral values over a period range
can be computed as shown in Equation 6.1 or, equivalently, by taking the (arithmetic)
mean of the natural logarithms of the spectral values. The scaled spectra shown in Figure
6.5(b) are re-plotted in Figure 6.5(c) using a logarithmic axis for the spectral values,
which makes it easier to observe differences in the spectra at longer periods and to see
that the Saavg values of the spectra are equal. For the conditions used to create the spectra
in Figure 6.5, it is expected that ground motions from Mw = 7.5 events will have greater
spectral values at moderate to long periods and smaller spectral values at short periods
relative to ground motions from Mw = 6.5 or 7.0 events for equal values of Saavg.
Chapter 6: Intensity Measures for Predicting Collapse 192

Figure 6.5: Response spectra predicted by BA08 for different magnitudes with Rjb =
12 km, Vs30 = 285 m/s, and unspecified fault type: (a) unscaled spectra; (b) spectra scaled
to a common value of Saavg (linear Sa(T) axis); (c) spectra scaled to a common value of
Saavg (logarithmic Sa(T) axis). The dashed vertical lines bracket the period range used to
compute Saavg.

As discussed in Chapter 5 and Section 6.3.2, the response spectrum is a signature of the
ground motion that reflects the duration and intensity of individual pulses or series of
pulses it contains (i.e., the frequency content). The frequency content of the ground
motion can significantly affect the response of a structure; however Saavg does not
directly account for relative differences in frequency content between ground motions
because it measures the average value of the spectrum. For example, two ground motions
can have the same Saavg value, such as the ground motions shown in Figure 6.5(c), but
one ground motion may have significant high frequency content (large spectral values at
short periods) while the other has significant moderate frequency content (large spectral
values at medium periods). Because Saavg does not directly account for relative
differences in the frequency content of ground motions that can affect the structural
response and because the frequency content of a ground motion is related to the
earthquake magnitude, it is not surprising that the p-values of the regressions between
Saavg,col and magnitude are somewhat close to the threshold between sufficiency and
insufficiency. These p-values do show, however, that there is some small influence of
earthquake magnitude on Saavg,col but that it is not statistically significant (at a 5%
significance level) for this case study.
Chapter 6: Intensity Measures for Predicting Collapse 193

The sufficiency of Saavg,col with respect to earthquake magnitude is further, and more
generally, supported by the case studies summarized in Table 6.2, which presents
statistics on the p-values of the linear relationship between ln(Saavg,col) and Mw for many
different structures briefly described earlier. As shown in Table 6.2, the mean p-value for
each structural type is significantly greater than 0.05. It is worth noting that only 53% of
the RC MRF structures show that Saavg,col is sufficient with respect to magnitude. This
may be related to the fact that the ground motions used for these structures cover a wider
range of magnitudes than those used for the other structures. One would expect that the
wider range of magnitudes would lead to larger differences in the relative frequency
content between the ground motions, making any relationship between Saavg,col and
magnitude more apparent. As previously stated, magnitude typically affects the frequency
content of the ground motion (i.e., the intensity and duration of the pulses in the ground
motion), which is not directly accounted for by Saavg and can affect the response of the
structure. The ground motions used for the generic MRF and generic shear wall structures
presented in Table 6.2 have a comparatively narrow magnitude range (Mw = 6.53 6.93
versus 6.50 7.62) so it is not surprising that their Saavg,col values demonstrate
significantly less dependency on magnitude than the RC MRF structures. In fact, all but 9
of the 648 generic MRF and generic shear wall structures show that Saavg,col was
sufficient with respect to magnitude at the 5% significance level.

Table 6.2: Summary of p-values for the relationship between Saavg,col and earthquake
magnitude.
No. of %
Structure Mw p-values
structures Sufficient*
Range Mean Range
Generic MRFs 6.53 6.93 396 98% 0.45 0.02 1.00
Generic Shear Walls 6.53 6.93 252 100% 0.63 0.08 1.00
RC MRFs 6.50 7.62 30 53% 0.21 2x10-5 0.90
*
Indicates percentage of structures with p-values of at least 0.05

To put these results into perspective, only 20% of the RC MRF structures are sufficient
with respect to magnitude when the IM is Sa(T1), currently the de-facto IM of choice,
compared to 53% of the structures when the IM is Saavg. While neither of these IMs
directly accounts for differences in the relative frequency content between ground
Chapter 6: Intensity Measures for Predicting Collapse 194

motions, Saavg is thought to be more sufficient with respect to magnitude because it


ensures that the average spectral value over a given period (frequency) range is equal
among the ground motions whereas Sa(T1) ensures that the spectral value at only a single
period (frequency) is equal among ground motions.

Interestingly, the RC MRF structures for which Sa(T1)col was sufficient or almost
sufficient with respect to magnitude were typically the same structures for which Saavg,col
was not sufficient with respect to magnitude and vice-versa. This is illustrated in Figure
6.6, which shows the p-value of the relationship between collapse intensity and
magnitude for each of the 30 RC MRF structures. The sufficiency of Sa(T1)col or Saavg,col
with respect to magnitude appears to be related to the number of stories and/or the first-
mode period of the structure as Structures 1 13 have between one and four stories and
T1 between 0.42 s and 1.16 s while Structures 14 30 have between 8 and 20 stories and
T1 between 1.57 s and 2.63 s. A possible reason why Saavg,col is not sufficient with respect
to magnitude for Structures 1 13 is that the dependence of spectral shape on magnitude
is typically most apparent in the short period range (the width and height of the spectral
peak), so the response of structures with shorter periods may be more likely to be affected
by the earthquake magnitude. As previously stated, Saavg cannot distinguish differences in
the relative frequency content between ground motions that can affect the structural
response.
Chapter 6: Intensity Measures for Predicting Collapse 195

Figure 6.6: p-values for the relationship between magnitude and ln(Saavg,col) or
ln(Sa(T1)col) for the RC MRF case study structures.

In summary, case studies of nearly 700 MRF and shear wall structures show that Saavg
(computed using a period range between 0.2T1 and 3T1) is, in general, sufficient with
respect magnitude. However, Saavg is not sufficient with respect to magnitude for the case
study RC MRF structures with T1 1.16 s, for which the ground motions span a notably
wider range of magnitudes (Mw from 6.50 to 7.62). For this subset of structures the
average slope of the linear regression between ln(Saavg,col) and Mw is -0.24 meaning that,
on average, the Saavg,col value found using a ground motion from a Mw = 6.50 event is
expected to be approximately 30% larger than the Saavg,col value found using a ground
motion from a Mw = 7.62 event. It should be noted, however, that the correlation between
ln(Saavg,col) and Mw is only moderate at best with an average correlation coefficient of =
0.38 and an average R2 value of 0.14 for these structures, indicating significant scatter in
the relationship.

6.3.3.2 Distance
Figure 6.7 shows the collapse intensities plotted against the source-to-site distance
measured using the Joyner-Boore distance (Rjb). As shown in Figure 6.7, a regression of
this data produces a slope of -0.002 and a p-value of 0.15, indicating that Saavg is
sufficient with respect to distance. It is not surprising that the trend between Saavg,col and
Chapter 6: Intensity Measures for Predicting Collapse 196

distance is weaker than the trend between Saavg,col and magnitude because distance
typically has much less effect on spectral shape than magnitude in GMPEs (Abrahamson
et al. 2008).

Figure 6.7: ln(Saavg,col) plotted against source-to-site distance with the ordinary linear
least squares regression of the data.

The sufficiency of Saavg with respect to distance is supported by other case studies as
shown in Table 6.3, which shows statistics on the p-values of the linear relationship
between ln(Saavg,col) and Rjb for many different case studies. As shown in Table 6.3, the
mean p-value for each structural type is significantly greater than 0.05.

Table 6.3: Summary of p-values for the relationship between Saavg,col and distance.
No. of %
Data set Rjb [km] p-values
structures Sufficient*
Range Mean Range
Generic MRFs 0.0 37.7 396 99% 0.53 0.03 1.00
Generic Shear Walls 0.0 37.7 252 100% 0.63 0.14 1.00
RC MRFs 0.9 74.2 30 93% 0.47 0.01 0.99
*
Indicates percentage of structures with p-values of at least 0.05

6.3.3.3 Spectral shape measures


This section investigates the sufficiency of collapse intensities measured by Saavg with
respect to spectral shape, specifically the parameters and SaRatio As previous results
showed that the values of Sa(T1)col were, in general, dependent on these parameters, it
Chapter 6: Intensity Measures for Predicting Collapse 197

was then desired to study whether measuring the collapse intensities using Saavg would
alleviate some of this dependence. Comparisons are made between the sufficiency of
Saavg and Sa(T1) with respect to these parameters.

6.3.3.3.1
Figure 6.8 shows the collapse intensities plotted with respect to the values for spectral
acceleration at T1 measured by the BA08 GMPE. As shown in Figure 6.8, a regression of
this data produces a slope of 0.017 and a p-value of 0.30, indicating that Saavg is sufficient
with respect to .

Figure 6.8: ln(Saavg,col) plotted against (T1) with the ordinary linear least squares
regression of the data.

Table 6.4 summarizes the p-values of the relationship between ln(Saavg,col) and for the
other case studies. Though the mean p-value for each structural type is above the 0.05
threshold for sufficiency, only slightly more than half of all structures considered have p-
values greater than 0.05. For perspective, only 3% of the generic MRF structures, only
1% of the generic shear wall structures and none of the RC MRF structures are sufficient
with respect to when the IM is Sa(T1), currently the most commonly used IM.
Chapter 6: Intensity Measures for Predicting Collapse 198

Table 6.4: Summary of p-values for the relationship between Saavg,col and .
No. of %
Data set p-values
structures Sufficient*
Mean Range
Generic MRFs 396 61% 0.16 4x10-4 0.98
Generic Shear Walls 252 41% 0.09 6x10-4 0.95
RC MRFs 30 43% 0.13 3x10-5 0.94
*
Indicates percentage of structures with p-values of at least 0.05

No trends were observed between structural properties such as number of stories or T1


and the sufficiency of Saavg,col with respect to ; however, a trend was observed with
respect to the period range used to compute Saavg. When using a period range of 0 to 3T1
(instead of 0.2T1 to 3T1) to compute Saavg, the percent of structures for which Saavg,col is
sufficient with respect to increases to 77% for the generic MRF structures and 81% for
the RC MRFs. On average, this change in period range does not affect the efficiency of
Saavg as an IM for these structures. As discussed in Chapter 7, the optimal period range
for maximum efficiency of Saavg is approximately 0 to 6T1 for the generic shear wall
structures. When Saavg is computed over a period range between 0 and 6T1 for the
generic shear wall structures, the percent of structures for which Saavg,col is sufficient with
respect to increases to 83%. Thus, it is believed that Saavg,col is, in general, sufficient
with respect to provided an appropriate period range is used to compute Saavg.

6.3.3.3.2 SaRatio
As discussed in Section 5.6, SaRatio is a spectral shape parameter that measures how
large Sa(T1) is relative to ordinates at neighboring periods and therefore is a direct
measure of spectral shape. When the collapse intensities measured by Saavg are plotted
against the SaRatio values using a log-log scale as shown in Figure 6.9, the slope is
approximately -0.08. The corresponding p-value is 0.04, which indicates that Saavg is
insufficient with respect to SaRatio at a 5% significance level; however, this value is very
close to the threshold between sufficient and insufficiency (a p-value of 0.05).
Chapter 6: Intensity Measures for Predicting Collapse 199

Figure 6.9: ln(Saavg,col) plotted against ln(SaRatio) with the ordinary linear least-squares
regression of the data.

Table 6.5 summarizes the p-values of the relationship between ln(Saavg,col) and
ln(SaRatio) for the other case studies. Though the mean p-value for each structural type is
above the 0.05 threshold for sufficiency, only approximately half of the generic shear
wall structures and half of the RC MRF structures considered have p-values greater than
0.05. For comparison, none of the 678 structures considered in Table 6.5 are sufficient
with respect to SaRatio when the IM is Sa(T1), currently the most commonly used (i.e.,
de-facto) IM of choice.

Table 6.5: Summary of p-values for the relationship between Saavg,col and SaRatio.
No. of %
Data set p-values
structures Sufficient*
Mean Range
Generic MRFs 396 88% 0.33 2x10-5 1.00
Generic Shear Walls 252 48% 0.14 2x10-5 0.98
RC MRFs 30 50% 0.20 7x10-9 0.88
*
Indicates percentage of structures with p-values of at least 0.05

Unlike the previous case in which computing Saavg over a different period range affected
the sufficiency of Saavg,col with respect to , using a different period range to compute
Saavg did not significantly affect the percentage of structures for which Saavg,col is
sufficient with respect to SaRatio. As previously discussed, Saavg does not distinguish
between the relative frequency content of records. The frequency content at T1 relative to
Chapter 6: Intensity Measures for Predicting Collapse 200

other periods as measured by SaRatio appears to affect the response of some structures
but not all. While on average Saavg,col is sufficient with respect to SaRatio, exceptions
exist as noted above.

6.3.4 Obtaining a hazard curve for Saavg


The previous case studies demonstrated that typically (1) Saavg is a significantly more
efficient IM for measuring the collapse intensity of a structure compared to Sa(T1) and (2)
Saavg is also more sufficient than Sa(T1) with respect to magnitude and spectral shape as
measured by and SaRatio. Research on and use of Saavg as an IM has increased in recent
years (e.g., Bianchini 2008; Bianchini et al. 2009; Bojrquez and Iervolino 2011;
Tsantaki et al. 2012; Shakib and Pirizadeh 2013); however, a significant obstacle to
adopting Saavg in practice is related to obtaining a hazard curve in terms of Saavg (SaAvg).
This is of concern whether Saavg is used for collapse risk assessment as done here or for
loss assessment due to non-collapse. Obtaining SaAvg is straightforward if one knows the
faulting rupture rates and probability distributions of magnitude and distance at a site
because existing GMPEs can be used to predict the probability of exceeding a given
value of Saavg conditioned on magnitude and distance (this is discussed later). The issue
of adopting Saavg in practice is that many structural engineers do not have access to the
data needed to compute a seismic hazard curve and/or rely on tools such as the USGS
Hazard Curve Application (USGS 2012a) or the OpenSHA Hazard Curve Calculator
(Field 2013), which currently provide Sa(T) but not SaAvg. It should be noted, however,
that currently Item 15 of OpenSHAs project goals includes implementing a GMPE for
Saavg, which would facilitate providing hazard curves in terms of this IM.

A hazard curve for a given IM is obtained through standard probabilistic seismic hazard
analysis (PSHA) calculations (see, e.g., Baker 2008) as follows

n mmax rmax
(IM > x ) = (M i > m min ) P(IM > x | m, r ) f M i ,Ri (m, r ) dr dm (6.7)
i =1 mmin 0
Chapter 6: Intensity Measures for Predicting Collapse 201

where (IM > x) denotes the mean annual frequency of the IM exceeding a value x, n is
the number of seismic sources considered, (Mi > mmin) is the mean annual frequency of
exceeding an event with magnitude mmin at source i, mmax and rmax are maximum
magnitude and distance, respectively, considered, P(IM > x| m,r) is the probability that
the IM exceeds a value x given values of magnitude and distance (which is provided by a
GMPE), and fMi,Ri(m,r) is the joint probability distribution function for magnitude and
radius.

If a GMPE for Saavg is not available then, as previously mentioned, existing GMPEs can
be used to estimate P(Saavg > x| m,r), permitting the computation of a hazard curve in
terms of Saavg. To explain how this is achieved it is helpful to look at the natural
logarithm of Saavg. Taking the natural logarithm of both sides of Equation 6.1 and
applying properties of logarithms yields the following expression for ln(Saavg)

N
( )
ln Sa avg (T1 , , T N ) =
1
N
ln(Sa(Ti )) (6.8)
i =1

Equation 6.8 shows that ln(Saavg) is computed as the sum of the ln(Sa(Ti)) terms
multiplied by a constant. Jayaram and Baker (2008) showed that the ln(Sa(Ti)) terms can
be assumed to be jointly normal, which as discussed by Baker and Cornell (2006a) and
Baker and Jayaram (2008) means that ln(Saavg) can then be assumed to be normally
distributed. Note that distribution of ln(Saavg) would not necessarily be normal if the
ln(Sa(Ti)) terms did not have a joint normal distribution.

Given that ln(Saavg) can be assumed to be normally distributed its probability distribution
is then completely defined by its expected (mean) value and variance (2), which are
computed as follows

N
[ ( )]
E ln Sa avg (T1 , , T N ) =
1
N
E[ln(Sa(Ti ))] (6.9)
i =1
Chapter 6: Intensity Measures for Predicting Collapse 202

N N
[ ]
2 ln (Sa avg (T1 , , T N )) =
1
ln(Sa(Ti )),ln (Sa (T j )) ln(Sa(Ti )) ln(Sa(Tj )) (6.10)
N2 i =1 j =1

where E[ ] denotes the expected value, 2[ ] indicates the variance, and ln(Sa(Ti)),ln(Sa(Tj)) is
the correlation coefficient between ln(Sa(Ti)) and ln(Sa(Tj)). Empirical equations have
been developed for estimating ln(Sa(Ti)),ln(Sa(Tj)). For example, Baker and Jayaram (2008)
developed an equation that can be used with four different GMPEs. It should be noted
that the E[ln(Sa(Ti))] terms shown in Equation 6.9 and the ln(Sa(Ti)) terms shown in
Equation 6.10 are readably available from existing GMPEs for Sa(Ti) and, although not
explicitly shown, are conditioned on the magnitude, distance, and other characteristics
such as site conditions and faulting mechanism. Therefore, E[ln(Saavg)] and 2lnSaAvg are
conditioned on the same magnitude, distance, and other site conditions used to compute
the E[ln(Sa(Ti))] and ln(Sa(Ti)) terms.

Computing the P(IM > x| m,r) term in Equation 6.7 for IM = Saavg is straightforward
using the information provided by Equations 6.9 and 6.10 and is shown as follows

( )
P Sa avg > x | m, r = 1
[ (
ln (x ) E ln Sa avg )]
(6.11)
ln SaAvg

where ( ) is the standard normal cumulative distribution function and E[ln(Saavg)] and
lnSaAvg are conditioned on the m and r values. Knowing the fault rupture rates and
distributions of magnitude and distance for each fault, a hazard curve for Saavg can be
then be computed with Equation 6.7.

Obtaining the values of E[ln(Saavg)] and lnSaAvg from Equations 6.9 and 6.10,
respectively, is referred here as the indirect approach as these values are computed using
E[ln(Sa(Ti))] and ln(Sa(Ti)), which are determined by regressions of the ln(Sa(Ti)) values
of individual ground motions. In a direct approach one would compute E[ln(Saavg)] and
lnSaAvg from a regression analysis directly on the ln(Saavg) values of individual ground
motions (i.e., developing a GMPE for ln(Saavg)). Reasons why the values of E[ln(Saavg)]
Chapter 6: Intensity Measures for Predicting Collapse 203

and/or lnSaAvg may different between the direct and indirect approaches are summarized
as follows: (1) different records may be used in a regression of ln(Sa(T)) versus a
regression of ln(Saavg) because of limitations associated with the maximum usable period
of some records; (2) equations for predicting E[ln(Sa(Ti))] with existing GMPEs do not
yield exact matches to the empirical data (i.e., residuals may differ from zero); and (3)
equations for predicting the correlation coefficients used Equation 6.10 also do not yield
exact matches to the empirical data.

Expanding on reason (1), if the maximum usable period of a record falls within the period
range used to compute ln(Saavg), the record would presumably be used to predict
ln(Sa(T)) values for periods less than the maximum usable period but would not be used
to compute ln(Saavg). Given the relatively large number of ground motions used by
typical GMPEs such as BA08 (typically at least 1,000 for moderate periods and several
hundred for a period of 10 s), differences in the ground motions used may not lead to
significant differences in the E[ln(Saavg)] and lnSaAvg values computed with each
approach. In reference to reasons (2) and (3), even though the equations used to predict
spectral ordinates and correlation coefficients do not produce exact matches to the
empirical data, Baker and Jayaram (2008) found good agreement between the correlation
coefficient values predicted by their equations and the empirical values, especially when
both periods of interest were less than five seconds.

The indirect approach is convenient as the terms in Equations 6.9 and 6.10 needed to
compute E[ln(Saavg)] and lnSaAvg, respectively, are readably available as they are
provided by existing GMPEs and approximate equations (in the case of the correlation
coefficients). On the other hand, the direct approach requires developing a new GMPE
using regressions of ln(Saavg). Furthermore, for the new GMPE to be as useful as existing
GMPEs such as the NGA GMPEs (Abrahamson et al. 2008) that provide the expected
values and variances of spectral ordinates at periods between T = 0.01 s and 10 s, the
regressions of ln(Saavg) would need to be performed for a myriad of period ranges.
Chapter 6: Intensity Measures for Predicting Collapse 204

6.4 Evaluation of different IMs for collapse risk assessment


This section evaluates the performance of different IMs, specifically Sa(T1), Saavg, and
peak ground velocity (PGV), with respect to their collapse risk predictions for a case
study. The discussion focuses on how the sufficiency and efficiency of different IMs
affects the collapse risk estimates as well as how the collapse risk estimate of a given IM
varies when different ground motions are used. The collapse risk estimate is quantified
using mean annual frequency of collapse (c). Sa(T1) is chosen as it is currently the most
widely used IM while Saavg is chosen because, as shown in the previous section, it
generally provides an improved prediction of structural response compared to Sa(T1).
PGV is also chosen herein as a proxy for the incremental velocity of pulses in the ground
motion as the velocity history of a ground motion is obtained by integrating the
acceleration history, and thus in many cases it may provide a relatively efficient
predictions of structural response. A case study by Bradley et al. (2010b) showed that the
dispersion of collapse intensities measured by Sa(T1) and PGV were comparable although
PGV produced lower collapse risk estimates. The latter was attributed primarily to the
ability of the GMPE to predict the value of PGV with less uncertainty compared to Sa(T1)
due to a smaller record-to-record variability.

The structure described in Section 3.5 is again used for evaluating the various IMs, and
the site is adapted from a hypothetical site used by Baker (2008, Section 1.4). The
seismicity at the site is dominated by a single fault located R = 10 km from the site. The
magnitude distribution is described by the bounded Gutenberg-Richter recurrence law,
which imposes a limit on the maximum earthquake (mmax) a source can produce. The
cumulative distribution function that describes FM(m), the probability of observing an
earthquake with a magnitude of less than or equal to m is

1 10 b ( m mmin )
FM (m ) = , m min < m < m max (6.12)
1 10 b ( mmax mmin )

where b is the relative ratio of small to large earthquakes in the region and mmin is the
minimum considered earthquake magnitude. For this site b = 1, mmin = 6, and mmax = 8.
Chapter 6: Intensity Measures for Predicting Collapse 205

The mean annual rupture rate of the fault is taken as (M > mmin) = 0.02. The fault rupture
is assumed to include the segment of the fault located nearest to the site so that Rjb, the
minimum horizontal distance from the site to the surface projection of the rupture area, is
constant and equal to 10 km. The average shear wave velocity at the site is Vs30 =
285 m/s.

The IMs considered include Sa(T1), Saavg and PGV, where T1 = 1.33 s and Saavg is
computed using a period range of 0.2T1 to 3T1 with a uniform period spacing of 0.01 s.
The hazard curves for each IM are computing using Equation 6.7 with the BA08 GMPE
used to describe the probability of exceeding a given intensity level conditioned on the
magnitude, distance, and site conditions. The correlation coefficients defined by Baker
and Jayaram (2008) are used to compute lnSaAvg using Equation 6.10. The seismic hazard
curves are shown in Figure 6.10(a) for Sa(T1) and Saavg and in Figure 6.10(b) for PGV.

6.4.1 Effects of efficiency and sufficiency


The MRCD 137 set of ground motions described in Section 3.3 and listed in Appendix A
are used here to analyze the response of the structure and obtain the collapse fragility
curve for each IM. The collapse fragility curves and c deaggregation curves are shown in
Figure 6.10(c) and Figure 6.10(e), respectively, for Sa(T1) and Saavg and in Figure 6.10(d)
and Figure 6.10(f), respectively, for PGV. Figure 6.10(a) shows that for a given intensity
level Sa(T1) has a greater mean annual frequency of exceedance than Saavg. It should be
noted that this does not necessarily lead to Sa(T1) producing a higher collapse risk
estimate than Saavg because the probability of collapse at a given intensity level is greater
for Saavg than Sa(T1) for nearly all intensity levels of interest as seen in Figure 6.10(c).
Chapter 6: Intensity Measures for Predicting Collapse 206

Figure 6.10: Collapse risk components for different IMs: Seismic hazard curves for
simplified site: (a) Sa(T1) and Saavg; (b) PGV. Collapse fragility curves for (c) Sa(T1),
Sa(T1) with or SaRatio adjustment, and Saavg; (d) PGV. c deaggregation curves for (e)
Sa(T1), Sa(T1) with or SaRatio adjustment and Saavg; (f) PGV. Note the different
horizontal scales between plots in the left and right columns as well as the different
vertical scales between plots (e) and (f).

A summary of relevant collapse risk metrics is provided for each IM in Table 6.6. The
mean annual frequency of collapse (c) and the probability of collapse in 50 years (Pc,50)
are computed for each IM using Equation 2.2 and Equation 2.5, respectively. As shown
in Table 6.6 each IM produces a different estimate of the collapse risk, with Sa(T1)
producing a Pc,50 estimate of approximately 1.0% versus estimates of approximately 0.3%
for both Saavg and PGV.
Chapter 6: Intensity Measures for Predicting Collapse 207

Table 6.6: Collapse risk metrics for different IMs with the MRCD 137 ground motion set.
Median Collapse
IM ln IM ln IM,BA08* c [10-4] Pc,50
Intensity
Sa(T1) 1.03 g 0.39 0.68 2.03 1.01%
Sa(T1) + 1.33 g 0.39 0.68 0.88 0.44%
Sa(T1) + SaRatio 1.40 g 0.39 0.68 0.74 0.37%
Saavg 0.70 g 0.22 0.59 0.59 0.29%
PGV 160 cm/s 0.33 0.58 0.50 0.25%
*
Note the BA08 in lnIM,BA08 indicates that this is the standard deviation from the BA08 GMPMGMPE
and not the standard deviation of the collapse intensities, which is denoted as lnIM.

One would like to know which IM gives the best estimate of the collapse risk. As
discussed in Section 2.3 the ground motions a structure will experience in its lifetime are
unknown, but one can estimate c using information about the types and frequencies of
seismic events that are likely to occur at the site and information about how the structure
is likely to respond to those events (i.e., whether or not they produce collapse). This is
accomplished as shown in Equation 2.4. The following discussion uses properties of the
IMs (e.g., efficiency, sufficiency, etc.) to argue why one IM gives a better estimate of c
than another.

Starting with Sa(T1), this IM is clearly not sufficient with respect to the spectral shape of
the ground motions used as discussed in Section 5.6. Given that Sa(T1) only considers a
single spectral value another step must be included to account for spectral shape when
using this IM to assess the collapse risk of structure. As discussed in Section 5.7, this can
be accomplished, for example, by using Sa(T1) as part of a vector IM with some measure
of spectral shape such as or SaRatio, carefully selecting records with regard to spectral
shape, or adjusting the results obtained with Sa(T1) when spectral shape is not directly
considered in record selection. Two approaches are considered here for modifying the
collapse risk estimate produced by Sa(T1): the adjustment discussed by Haselton et al.
(2011) and the SaRatio adjustment, which is discussed in Section 5.7 and is an adaptation
of the method discussed by Haselton et al.

As shown by several recent research studies (e.g., Baker and Cornell 2005b; Zareian and
Krawinkler 2007a; Haselton et al. 2011) records with positive values generally tend to
Chapter 6: Intensity Measures for Predicting Collapse 208

produce smaller structural responses (e.g., smaller interstory drift ratios) than records
with neutral or negative values for a given value of Sa(T1). With the exception of some
regions in eastern North America, the target values obtained from PSHA
deaggregations are almost universally positive in active seismic regions at long return
periods of engineering interest (Burks and Baker 2012). This means that for most sites
there is a strong likelihood of overestimating the collapse risk when using Sa(T1) as the
IM unless the ground motions are selected considering or some adjustment (i.e., a
reduction) is made to the results. This has been demonstrated by several researchers
including Baker and Cornell (2006a), Haselton and Deierlein (2007, Chapter 6) and
Haselton et al. (2011).

The MRCD 137 ground motion set used for the case study is approximately neutral (the
mean value of the ground motions is (T1) = 0.13). A PSHA deaggregation by
magnitude and at Sa(T1) = 0.72 g is shown in Figure 6.11 and is used to determine the
target value. This intensity level is chosen because it is at the peak of the c
deaggregation (Figure 6.10(e)), meaning it has the most significant contribution to the
collapse risk. Note that this intensity is very close to the 2/50 hazard level, which is
associated with Sa(T1) = 0.68 g. The PSHA deaggregation for the site presented in Figure
6.11 shows that a Mw = 6.7 event, which is associated with (T1) = 1.8, has the greatest
contribution to the seismic hazard at this intensity level. Given that the mean value of
the ground motion set ((T1) = 0.13) is significantly lower than the target value ((T1) =
1.8), the collapse risk computed using Sa(T1) has likely been overestimated compared to
the risk computed using ground motions with values near the target value. To estimate
the median collapse intensity that would be computed using records with values near
the target value, the adjustment discussed by Haselton et al. (2011) is used. The
regression of ln(Sa(T1)col) versus presented in Figure 5.29 shows that at the target value
of (T1) = 1.8 the expected median collapse intensity is Sa(T1)col = 1.33 g (up from 1.03 g,
see Table 6.6). Haselton et al. note that adjusting the median collapse intensity in this
manner also reduces the dispersion of the collapse intensities, and they provide an
equation for calculating this reduced dispersion. The reduction in dispersion is a function
of the slope and residuals from the regression as well as the standard deviation of the
Chapter 6: Intensity Measures for Predicting Collapse 209

values from the PSHA deaggregation. Haselton et al. found that the reduction in
dispersion was only on the order of 10% and had only a minor effect on the collapse risk.
For this reason they proposed simply using the dispersion calculated from the original
(i.e., unadjusted) collapse intensities.

Figure 6.11: PSHA deaggregation by magnitude with associated (T1) values for
Sa(T1) = 0.72 g, the intensity at the peak of the c deaggregation.

The -adjusted estimate of the collapse fragility curve for Sa(T1) (denoted as Sa(T1)+)
is shown in Figure 6.10(c) and is used to compute c. The resulting c deaggregation is
presented in Figure 6.10(e). The collapse risk measured by Sa(T1)+ is Pc,50 = 0.44% as
shown in Table 6.6. This is approximately a 60% reduction in Pc,50 compared to the value
calculated for Sa(T1) without considering ; however, this adjusted value of Pc,50 is still
approximately 1.5 or 1.8 times greater than the Pc,50 computed using Saavg or PGV,
respectively.

Another method of adjusting the collapse fragility curve obtained using Sa(T1) as the IM
is to use the spectral shape parameter SaRatio. This approach is described in Section
5.7.2.1. The target value of SaRatio is estimated to be 2.06, which is obtained from a
conditional mean spectrum (CMS) (see Baker and Cornell 2006a; Lin et al. 2013b) of the
event with the largest contribution to the seismic hazard as determined by Figure 6.11.
The CMS is computed using the BA08 GMPE with the following parameter values: Mw =
6.7, Rjb = 10 km, (T1) = 1.8, Vs30 = 285 m/s, and an unspecified fault type. Correlation
Chapter 6: Intensity Measures for Predicting Collapse 210

coefficients between spectral values at different periods are computed using the equations
developed by Baker and Jayaram (2008). Following the approach described in Section
5.7.2.1 the modified median collapse intensity is computed from the regression between
ln(Sa(T1)col) and ln(SaRatio) shown in Figure 5.28 using the target value of SaRatio. This
results in a modified median collapse intensity of Sa(T1)col = 1.40 g (up from 1.03 g, see
Table 6.6).

The SaRatio-adjusted estimate of the collapse fragility curve for Sa(T1) (denoted as
Sa(T1)+SaRatio) is shown in Figure 6.10(c) and is used to compute c. The resulting c
deaggregation is presented in Figure 6.10(e). The collapse risk measured by
Sa(T1)+SaRatio is Pc,50 = 0.37% as shown in Table 6.6. This is approximately a 60%
reduction in Pc,50 compared to the value calculated for Sa(T1) without considering
SaRatio; however, although lower than the value computed with the -adjusted Sa(T1),
this SaRatio-adjusted value of Pc,50 is still approximately 1.3 or 1.5 times greater than the
Pc,50 computed using Saavg or PGV, respectively.

It is believed that a primary reason why Sa(T1) (with or without the spectral shape
adjustments) produces a higher estimate of the collapse risk is because it is less correlated
with collapse intensities, and therefore it is less efficient than either Saavg or PGV. The
efficiencies of the IMs with respect to the median collapse intensity are compared in
Table 6.6, and it is shown that Sa(T1) is the least efficient IM because it has the largest
value of lnIM. Furthermore, Sa(T1) is also associated with the largest variability from the
GMPE perspective (i.e., the variability in the value of the IM for a given magnitude and
distance), which also contributes to the higher collapse risk estimates for this IM and will
be discussed later. Saavg, on the other hand, combines the smallest dispersion in the
collapse intensities with a smaller variability in the IM from the GMPE perspective,
which as discussed later leads to more reliable estimate of the collapse risk.

For a given median collapse intensity, the larger the dispersion the larger the values of
P(C|IM) in the lower half of the collapse fragility curve. As discussed in Sections 3.5 and
3.6, when the IM is Sa(T1) the most significant contributions to the collapse risk typically
Chapter 6: Intensity Measures for Predicting Collapse 211

come from intensities in the lower half of the collapse fragility curve. This is the case for
this particular case study as shown by the c deaggregation in Figure 6.10(e), which is
repeated in Figure 6.12(e), and means that the increased dispersion seen for Sa(T1) is an
important reason why Sa(T1) produces a larger collapse risk estimate than the other IMs.

For example, consider what would happen if Sa(T1) were as efficient as Saavg. Figure
6.12(c) illustrates the effect on the collapse fragility curve if lnSa(T1) = 0.22, the value of
lnSaAvg, rather than lnSa(T1) = 0.39, the value observed in the case study: the reduced
dispersion results in significantly lower values of P(C|IM) for values less than the median
collapse intensity, especially for intensities between 0.5 g and 1.0 g. The impact on the
collapse risk estimate is evident when examining the c deaggregation curve shown in
Figure 6.12(e) as the reduced dispersion significantly decreases the collapse risk
contribution from intensities less than 1.0 g, which are responsible for the majority of the
collapse risk observed in the case study. The collapse risk computed using the reduced
dispersion of lnSa(T1) = 0.22 is Pc,50 = 0.61%, which is a significant reduction from the
original value of Pc,50 = 1.01% (Table 6.6). This example is provided merely to illustrate
that the efficiency of an IM affects the collapse risk estimate and that for Sa(T1) an
increase in the efficiency (which corresponds to a decrease in lnSa(T1)) leads to a
reduction in the collapse risk estimate in most cases of engineering interest 2.

2
Note that in some cases it is possible for an increase in the efficiency of an IM (a decrease in lnIM) to lead
to an increase in the collapse risk. This would usually occur in a situation where c is dominated by
intensities in the upper half of the collapse fragility curve because decreasing lnIM while holding the
median constant increases P(C|IM) for the upper half of the fragility curve. Note that such a situation is,
however, rare in engineering practice, at least when the IM is Sa(T1). For example, for the case study
discussed in this section if the median collapse intensity were Sa(T1) = 0.1 g, a value of lnSa(T1) = 0.4 would
produce a slightly smaller value of Pc,50 than if lnSa(T1) = 0.2 (48.2% versus 49.0%). Given the high
probabilities of collapse, hopefully such a structure would not be built.
Chapter 6: Intensity Measures for Predicting Collapse 212

Figure 6.12: Collapse risk components for Sa(T1) illustrating the effects of improved
efficiency and improved GMPE: Seismic hazard curves for simplified site in (a) and (b);
Collapse fragility curves in (c) and (d); c deaggregation curves in (e) and (f). Plots (a),
(c) and (e) illustrate the effect of improved efficiency while plots (b), (d) and (f) illustrate
the effect of improved GMPE prediction.

Another important factor that contributes to the typically greater collapse risk estimates
given by Sa(T1) is related to the seismic hazard. The expected value and standard
deviation (lnIM,GMPE) of an IM are given by a GMPE and are used to calculate the
probability of exceeding a given intensity level as shown in Equation 6.11. This
information is then used to compute the seismic hazard curve as shown in Equation 6.7.
The standard deviations indicate how well the IM values observed from recorded ground
motions match the GMPEs prediction of the expected value and as such indicate, loosely
speaking, how easy or difficult it is to predict the value of the IM. As noted by Bradley et
al. (2010b), a large uncertainty in the GMPE will significantly increase the hazard
Chapter 6: Intensity Measures for Predicting Collapse 213

especially at low values of IM which are of primary interest in earthquake resistant


design.

The BA08 GMPE is used for all IMs considered in this case study and produces the
standard deviations lnIM,BA08 shown in Table 6.6 assuming an unspecified fault type.
Perhaps it is not surprising that Saavg has a smaller value of lnIM,BA08 than Sa(T1) as one
might expect that the process of averaging spectral ordinates over a range of periods
reduces extreme values. The mathematical basis for why lnSaAvg,BA08 is smaller than
lnSa(T1),BA08 can be understood by examining the terms in Equation 6.10, which shows that
lnSaAvg,BA08 is a function of the lnSa(T),BA08 values for periods used to compute Saavg and
the correlation coefficients for spectral values at different periods. Figure 6.13(a) shows
the former while Figure 6.13(b) shows the latter. For the case study Saavg is computed
between periods of 0.27 s and 3.99 s (which is the period range between 0.2T1 and 3T1,
where T1 = 1.33 s), which means this period range has over twice as many periods greater
than T1 than less than T1. Although periods greater than T1 have greater values of
lnSa(T),BA08 compared to the value at T1 as seen in Figure 6.13(a), the spectral values are
not perfectly correlated as seen in Figure 6.13(b). The reduction in the correlation of
spectral ordinates as one departs from T1 then, in practically all cases, leads to a reduction
in lnSaAvg,BA08 compared to the value of lnSa(T1),BA08.

Figure 6.13: Components influencing the value of lnSaAvg,BA08: (a) values of lnSa(T),BA08;
and (b) contour of correlation coefficients between spectral values at different periods.

To test the impact of the relatively higher value of lnSa(T1),BA08 compared to the other IMs,
the hazard curve for Sa(T1) is recomputed using a reduced standard deviation of
Chapter 6: Intensity Measures for Predicting Collapse 214

lnSa(T1),BA08 = 0.58, which is comparable to the value of lnIM,BA08 for Saavg and PGV, and
the collapse risk is recalculated. Figure 6.12(b) compares the original hazard curve for
Sa(T1), which is based on lnSa(T1),BA08 = 0.68, to the hazard curve computed using the
reduced standard deviation of lnSa(T1),BA08 = 0.58. Reducing the standard deviation of the
GMPE leads to a reduction in the seismic hazard, especially as the level of intensity
increases. Figure 6.12(f) compares the c deaggregation curves obtained using the
different hazard curves in Figure 6.12(b) and shows that the reduction in collapse risk
contribution is most significant for intensities between approximately 0.5 g and 1.5 g.
Though significant differences exist in the hazard curves for intensities greater than 1.5 g,
the likelihood of observing these larger intensities is so small compared to that of the
smaller intensities (i.e., the slope of the seismic hazard curve is relatively small) that the
larger intensities have relatively little impact on the collapse risk. The collapse risk
computed using the reduced GMPE standard deviation of lnSa(T1),BA08 = 0.58 is Pc,50 =
0.56%, which is a significant reduction from the original value of Pc,50 = 1.01% (Table
6.6) and demonstrates that the larger value of lnIM,BA08 for Sa(T1) relative to the other IMs
also plays a role in why Sa(T1) produces higher collapse risk estimates than the other
IMs.

As shown in Table 6.6 the collapse risk estimates given by Saavg and PGV are very
similar, with Saavg estimating c and Pc,50 values approximately 15% greater (in relative
terms) than the PGV estimates. Saavg is significantly more efficient than PGV as
illustrated by the 33% reduction in dispersion when comparing lnSaAvg to lnPGV (i.e., lnIM
= 0.22 versus 0.33). As previously discussed, with all else being equal one would
generally expect that the more efficient IM would produce smaller estimates of the
collapse risk because of the reduced uncertainty (reduced dispersion) in the collapse
intensities. However, this is not the case here as PGV (the less efficient IM) leads to
smaller collapse risk estimates than Saavg (the more efficient IM). This is because, even
though the collapse intensities measured by PGV have more variability about the median
than those measured by Saavg, once the discrete P(C|IM) values are multiplied by the
difference in the seismic hazard between the IM levels the terms (technically, the sum of
these terms) are smaller for PGV than Saavg. This is illustrated in Figure 6.14, which
Chapter 6: Intensity Measures for Predicting Collapse 215

shows P(C|IM) plotted as a function of the mean annual frequency of exceedance of the
IM level, IM, for both Saavg and PGV. As seen in Figure 6.14, for mean annual
frequencies smaller than 10-4 (which are typically of interest for collapse assessment) the
P(C|IM) at a given value of IM is smaller for PGV than Saavg

Figure 6.14: The probability of collapse conditioned on intensity level, P(C|IM), plotted
as a function of IM, the mean annual frequency of exceedance of the IM level for Saavg
and PGV.

6.4.2 Effect of ground motion selection


Previous results showed that both Saavg and PGV gave relatively comparable estimates of
the collapse risk. To test the sensitivity of this observation to the particular ground
motions used, the collapse fragility curve was recomputed using different sets of ground
motions and the collapse risk estimates were calculated for each set. Ideally, a good IM
will be more robust by being less sensitive to the ground motion set, that is, a good IM
will be one that gives relatively stable estimates of the collapse fragility and collapse risk
no matter what ground motions are used (i.e., smaller variability in the various estimates
computed with different ground motion sets) because it is hoped that the IM is sufficient
with respect to magnitude, distance, and other properties of the ground motions that
affect the structural response.
Chapter 6: Intensity Measures for Predicting Collapse 216

6.4.2.1 Description of ground motion sets


The first ground motion set considered was the expanded Large Magnitude-Short
Distance set compiled and used by Medina and Krawinkler (2003, Table 3.6) and is
referred to as LMSR-N. This set consists of 40 recordings where each record is from a
separate station (i.e., one random horizontal component is used from each station) from
events with magnitudes of Mw = 6.53 6.93 and distances of Rjb = 0.0 km 37.7 km.
Additional details about the records in this set are provided by Medina and Krawinkler
(2003, Section 3.3).

The Set One ground motions are compiled and documented by Haselton and Deierlein
(2007, Table 3-7). Set One consists of 39 pairs of recordings, each with two horizontal
components, from events with magnitudes of Mw = 6.50 7.62 and distances of Rjb =
0.9 km 74.2 km. Additional details about the records in this set are provided by
Haselton and Deierlein (2007, Appendix 3B).

The Set #1A, FN/FP ground motions were compiled and used by Baker et al. (Baker et
al. 2011). Unlike the previous ground motion sets, these ground motions were selected
with respect to spectral shape. For this ground motion set, records were selected so that
the spectral ordinates between periods of 0 and 5 s matched the means and standard
deviations predicted by the BA08 GMPE for a magnitude 7 strike-slip earthquake at a
distance of 10 km from a site with an average shear wave velocity of Vs30 = 250 m/s.
Forty recordings, each with two horizontal components, were selected and the resulting
range of magnitudes and distances are Mw = 6.06 7.90 and Rjb = 0 km 40.4 km. All
records were rotated to the fault-normal and fault-parallel (FN/FP) components.
Additional details on this ground motion set are provided by Baker et al. (2011, Section
3.1). For comparison purposes the collapse risk is also evaluated using the same ground
motions but this time they are not rotated to FN/FP components (i.e., they are the
orientations provided by the PEER NGA database (Chiou et al. 2008)). This record set is
denoted Set #1A, unrotated.
Chapter 6: Intensity Measures for Predicting Collapse 217

The final ground motion are selected to match a conditional spectrum (CS) and are
denoted as the CS, (T1) = 1.7 and CS, (T1) = 1.8 sets, respectively. As discussed by
Lin et al. (2013a; 2013b), a CS describes the full distribution of the spectrum at all
periods (i.e., it captures both the expected spectral values and the variability of those
values) and is constructed using the properties of a main causal or target event (e.g.,
magnitude, distance, , fault type, etc.) with a GMPE or multiple causal events with
multiple GMPEs. Only a single target event and GMPE are used for the CS sets
constructed here. The target event is obtained from a PSHA deaggregation at Sa(T1) =
0.68 g, which corresponds to the 2/50 hazard level. This hazard level is chosen as it is
expected to have a significant contribution to the collapse risk. The magnitude with the
largest contribution to the hazard is Mw = 6.7, which corresponds to (T1) = 1.7. The CS is
constructed using these Mw and (T1) values with the other site parameters (i.e., Rjb = 10
km, Vs30 = 285 m/s, and an unspecified fault type) and the BA08 GMPE. Records are
selected to match the CS using publicly available software described by Jayaram et al.
(2011). Forty recordings, each with two horizontal components, are selected from the
PEER NGA database to match this CS and are denoted as the CS, (T1) = 1.7 set. A
CS, (T1) = 1.8 set is also constructed in the same manner with the same parameter
values except that the (T1) value is 1.8. This was done to be consistent with the target
(T1) value of the MRCD 137 set as shown in Figure 6.11 and to be closer to the target
(T1) values of other ground motion sets, which are discussed later. Additional
information about the CS, (T1) = 1.7 and CS, (T1) = 1.8 sets including the NGA
identification numbers is provided in Appendix C.

It should be noted that although the record selection methodology and the values of many
target parameters (e.g., magnitude, distance, etc.) are similar for the Set #1A and the CS
sets, there is an important difference between these two sets: the CS sets were
conditioned on non-zero values of (T1), which affects the target spectrum due to
correlations between values at different periods. The positive (T1) values used to
construct the target spectra for the CS sets imply that these spectra typically have a peak
at or near T1. In contrast, the target spectrum used to select the Set #1A ground motions is
Chapter 6: Intensity Measures for Predicting Collapse 218

based on the median prediction from the GMPE, which implies that the values are zero
for all periods so the spectral shape characteristics of the two sets are very different.

As discussed in Chapter 5, the spectral shapes of records can significantly affect the
structural response, especially at large levels of nonlinearity. Comparisons of the
(geometric) mean spectrum of each ground motion set are presented in Figure 6.15,
where all records are scaled to a common value of Sa(T1) in Figure 6.15(a), a common
value of Saavg in Figure 6.15(b), and a common value of PGV in Figure 6.15(c). To
compute the geometric mean spectrum of a given record set for a given IM, all records as
scaled to the same IM value and the geometric mean spectral value is computed at all
periods. Figure 6.15 shows that the overall (i.e., mean) relative frequency content differs
among ground motion sets. For example, the MRCD 137 and both Set #1A ground
motion sets generally have higher spectral ordinates compared to the other ground motion
sets between periods of 3 s and 4 s while the LMSR-N and Set One sets generally have
higher spectral ordinates compared to the other ground motion sets at short periods,
regardless of which IM is used to scale the records. The CS sets exhibit peaked spectra
near T1 and tend to have higher spectral ordinates near T1 compared to the other ground
motion sets when scaled to a common value of Saavg or PGV.
Chapter 6: Intensity Measures for Predicting Collapse 219

Figure 6.15: Geometric mean spectrum of different ground motion sets scaled to a
common value of (a) Sa(T1); (b) Saavg; and (c) PGV. Note the vertical line in (a) is drawn
at T1 = 1.33 s and the horizontal line in (b) denotes the common value of Saavg over the
period range between 0.2T1 and 3T1.

6.4.2.2 Collapse risk comparisons


The collapse risk computed by each IM for the different ground motion sets are
summarized in Figure 6.16. The -adjusted and SaRatio-adjusted collapse risk results for
Sa(T1) are also presented in Figure 6.16 and are denoted as Sa(T1)+ and
Sa(T1)+SaRatio, respectively. A PSHA deaggregation at the intensity associated with
the peak of the c deaggregation is used to determine the target and SaRatio values
using the procedures previously described. This intensity varies between ground motion
sets (from Sa(T1) = 0.70 g to 0.87 g), which results in the target values of these
parameters varying between the ground motion sets (from (T1) = 1.8 to 2.1 and from
SaRatio = 2.1 to 2.2). All PSHA deaggregations show that the magnitude with the
greatest contribution to the seismic hazard is Mw = 6.7.
Chapter 6: Intensity Measures for Predicting Collapse 220

Figure 6.16(a) presents the median collapse intensities for each IM. The median collapse
intensities for the IMs associated with spectral accelerations are reported in units of g
while the values for PGV are reported in cm/s. As previously discussed, it is desirable
that the median collapse intensity of a given IM remain relatively stable for different
ground motion sets because this indicates that the results are likely not affected by
properties of the ground motions such as magnitude, distance, , or other properties that
differ between the ground motion sets, and therefore the results are more reliable as they
are less affected by decisions made when selecting ground motions. Figure 6.16(a) shows
that the median collapse intensities measured by Saavg are very robust and relatively
stable as they only vary between Saavg = 0.70 g and 0.77 g (a 5% variation with respect to
the value in the middle of this range of intensities). The median collapse intensities
measured by Sa(T1)+, on the other hand, vary considerably more as the values range
from Sa(T1) = 1.33 g to 1.72 g (a 13% variation with respect to the value in the middle of
this range of intensities).
Chapter 6: Intensity Measures for Predicting Collapse 221

Figure 6.16: Summary of collapse risk results for different IMs and different ground
motion sets: (a) median collapse intensity; (b) standard deviation of collapse intensities,
lnIM; (c) mean annual frequency of collapse, c, and probability of collapse in 50 years,
Pc,50. Note in plot (a) the spectral acceleration-based IMs to the left of the dashed vertical
line are in units of g while PGV, which is to the right of the dashed vertical line, is in
units of cm/s.

The dispersions (lnIM) on the median collapse intensity are presented in Figure 6.16(b)
and show that for all ground motion sets Saavg is the most efficient IM, with an average
dispersion value of lnSaAvg = 0.25. PGV has an average dispersion value of lnPGV = 0.33,
and Sa(T1) is the least efficient IM with an average dispersion value of lnSa(T1) = 0.39.
Note that the dispersion of collapse intensities computed using the Sa(T1)+ or
Sa(T1)+SaRatio methods were not directly computed as the reduction in dispersion
compared to Sa(T1) alone is expected to be small and were instead taken as equal to the
dispersion of collapse intensities based on Sa(T1) alone. As previously discussed the
Chapter 6: Intensity Measures for Predicting Collapse 222

efficiency of an IM affects the collapse risk results, and generally for a given IM an
increase in efficiency (i.e., a decrease in dispersion) leads to a reduced collapse risk
estimate. This occurs because a decrease in dispersion reduces the probabilities of
collapse at intensities less than the median, which typically contribute most significantly
to c. It was also previously discussed that just because one IM is more efficient than
another does not necessarily mean it will produce a lower collapse risk estimate. This is
because the mean annual frequencies of different IMs differ, and the collapse fragility
needs to be combined with the seismic hazard as done when computing c to determine
the collapse risk.

The values of c and Pc,50 for each IM are presented in Figure 6.16(c). Discussion herein
will focus on the Pc,50 values as they are simply a transform of the c values as shown by
Equation 2.5. Figure 6.16(c) shows that collapse risk estimates are most stable for Saavg,
which has Pc,50 values ranging from 0.22% to 0.33% (a 20% variation with respect to the
value in the middle of this range). This indicates that compared to the other IMs the
collapse risk estimated using Saavg is more robust and relatively insensitive to the
particular set of ground motions used, which is a very desirable property. This relatively
narrow range of collapse risk estimates is obtained using records sets with very different
characteristics and yet they lead to very similar collapse risk estimates. For example
Figure 6.15(b) shows that when scaled to the same Saavg value records in the MRCD 137
set have, on average, significantly larger spectral ordinates than the LMSR-N set between
periods of 3 s and 4 s while between periods of 0 and 1 s the trend is reversed. Despite
differences in relative frequency content that Saavg does not capture, the collapse risk
estimates using these two sets are very similar as the Pc,50 values are 0.29% and 0.23%
for the MRCD 137 and LMSR-N sets, respectively.

Sa(T1)+ produces collapse risk estimates ranging from Pc,50 = 0.14% to 0.44% (a
variation of approximately 50% with respect to the value in the middle of this range),
which indicates that even after correcting for the different values between the ground
motion sets the collapse risk computed by this IM is still sensitive to the records used in
the evaluation. The collapse risk estimates obtained using Sa(T1)+SaRatio show less
Chapter 6: Intensity Measures for Predicting Collapse 223

sensitivity to the ground motion sets than those obtained using Sa(T1)+, with estimates
of the former ranging from Pc,50 = 0.22% to 0.40% (a variation of approximately 30%
with respect to the value in the middle of this range). This is due to less variation in the
median collapse intensity estimates as this is the only parameter affecting the collapse
risk estimate that differs between the two IMs. The reduced variation in the median
collapse intensity estimates is thought to be the result of significantly greater correlation
of the Sa(T1)col values with SaRatio as opposed to , which was demonstrated in
Section 5.6.

The collapse risk estimates using PGV vary from Pc,50 = 0.14% to 0.52%, which is larger
than the range using Sa(T1)+, and indicate that this IM is also sensitive to the records
used. Knowing the value of PGV does not provide information about the response
spectrum or its shape, which influences structural response, so it is not surprising that the
collapse risk estimates obtained with different ground motions vary so widely for PGV.
For example a comparison of the mean response spectra for the different ground motion
sets scaled to the same PGV in Figure 6.15(c) with the collapse risk estimates in Figure
6.16(c) shows that the CS sets have significantly larger spectral ordinates near T1 and
produce higher collapse risk estimates than the other sets.

The results discussed here suggest that Saavg is a good IM to use for collapse risk
assessment as compared to the other IMs examined it is a better (more robust and more
efficient) predictor of structural response, is predicted with less or comparable dispersion
from a GMPE perspective, and produces stable collapse risk estimates using different
ground motion sets. The latter is attributed to its sufficiency with respect to magnitude,
distance, , and other ground motion properties that can affect the structural response.

6.5 Summary
This chapter evaluated the performance of different IMs as it related to predicting the
collapse risk of a structure. Properties of a good IM include efficiency (a measure of the
amount of uncertainty in the predicted structural response), sufficiency (that the structural
response is not affected by properties of the ground motion not measured by the IM), and
Chapter 6: Intensity Measures for Predicting Collapse 224

the ability to compute hazard information about the IM as well as the relative ease with
which the information is obtained. Related to efficiency and sufficiency, the ability of the
IM to produce reliable structural response estimates and robust risk estimates that are not
sensitive to the ground motions used in analysis is also desirable.

Sa(T1) has commonly been used as an IM for predicting nonlinear structural response
including collapse because, unlike IMs such as PGA or PGV, it provides information
about how the ground motion affects the structure (specifically, an elastic SDOF system).
A review of past research on IMs showed that including information about either the
spectral shape of the ground motion indirectly through a parameter like or, even better,
directly by incorporating information about the spectral ordinates at periods other than or
in addition to the fundamental period improved response predictions compared to using
Sa(T1) alone.

Chapter 5 investigated how and why the spectral shape of a record affects the structural
response, and it was seen that spectral acceleration averaged over a period range (Saavg)
could give a good prediction of which records were more damaging when scaled to the
same value of Sa(T1). Saavg was used as an IM in this chapter and was compared to other
IMs. Saavg, which is calculated as the geometric mean of spectral acceleration over a
period range, was shown to be significantly more efficient than Sa(T1) at predicting the
collapse intensity of a ground motion. This was attributed to several factors, including:
(1) that Sa(T1) does not consider the (elastic) response at other periods which might be
important in estimating nonlinear and MDOF structural response; (2) that a change in the
displacement and velocity of the system when it is hit by significant ground motion
pulses can significantly alter the peak elastic response, and even though Saavg is also
based on elastic ordinates, the influence of the initial conditions at the arrival of
significant pulses is minimized by averaging the response over a range of periods; and (3)
by taking an average spectral value over a range of periods, Saavg accounts directly for the
spectral shape and indirectly for the underlying pulses in the ground motion that produce
the spectral shape and cause damage.
Chapter 6: Intensity Measures for Predicting Collapse 225

The sufficiency of Saavg was evaluated with respect to magnitude, distance, and spectral
shapes measures and SaRatio using case studies of nearly 700 structures including
generic MRF, generic shear wall, and RC MRF structures. It was shown that Saavg is
generally sufficient with respect to all these measures, especially distance. A particular
subset of structures (RC MRF structures with T1 1.16 s, for which the ground motions
span a particularly wide range of magnitudes (Mw from 6.50 to 7.62)) was identified for
which Saavg was not sufficient with respect to magnitude. Considering all the RC MRF
structures, however, Saavg was generally significantly more efficient with respect to
magnitude than Sa(T1). Whereas virtually none of the 650+ structures were sufficient
with respect to or SaRatio when the IM was Sa(T1), the majority of structures were
sufficient with respect to these spectral shape measures when the IM was Saavg, especially
if the period range used to compute Saavg resulted in optimum or near optimum efficiency
levels.

Obtaining a seismic hazard curve for Saavg was discussed, and it was demonstrated how
one could be obtained using existing GMPEs that predict spectral acceleration values and
readily available equations to estimate the correlation between spectral ordinates at
different periods. It was noted that though seismic hazard curves in terms of Sa(T1) are
relatively widely available, this is not the case for hazard curves in terms of Saavg and is a
significant factor preventing the widespread use of the IM in practice.

This chapter concluded by evaluating the performance of different IMs in predicting the
collapse risk of a structure. Sa(T1), Saavg, and PGV were evaluated using a case study
structure with seven different ground motion sets. The collapse risk results computed
using Sa(T1) varied noticeably between the ground motion sets (i.e., a probability of
collapse in 50 years from Pc,50 = 0.14% to 0.44% or from Pc,50 = 0.22% to 0.40%,
depending whether or SaRatio was used to adjust the results for spectral shape). This
variance indicates that Sa(T1) is not completely sufficient because properties of the
ground motions other than Sa(T1) affected the structural response and lead to collapse
risk estimates which can be very sensitive to ground motion selection. A similar
Chapter 6: Intensity Measures for Predicting Collapse 226

conclusion was drawn for PGV as it also produced a relatively large range of collapse
risk estimates for the different ground motion sets (from Pc,50 = 0.14% to 0.52%).

Based on the case study results, Saavg was shown to have the best performance as
compared to the other IMs it was a better (more efficient and more robust) predictor of
structural response, was predicted with less or comparable dispersion from a GMPE
perspective, and produced relatively stable collapse risk estimates using different ground
motion sets. The latter is attributed to its sufficiency with respect to magnitude, distance,
, and other ground motion properties that can affect the structural response. This is
promising and suggests that the careful record selection and/or modification of structural
response results required to obtain a good estimate of the collapse risk when using Sa(T1)
as the IM may not be required if Saavg is the IM. Future work considering additional
structures will be necessary to confirm this.
Chapter 7

Average Spectral Acceleration (Saavg): Optimal Period Ranges


for Various Structures for Maximum Efficiency

7.1 Overview
This chapter focuses exclusively on Saavg and explores how the range, number, and
spacing (i.e., arithmetic versus logarithmic) of periods used to compute Saavg influences
the effectiveness of this intensity measure (IM). Particular focus is given to the impact
these variables have on the efficiency of the IM, though a discussion of their impact on
collapse risk assessment estimates is also included.

The period ranges that maximize the efficiency of Saavg are investigated for a variety of
structures including generic SDOF systems, generic moment-resisting frame (MRF)
systems, generic shear wall systems, and reinforced concrete (RC) MRF systems.
Recommendations about the optimum period ranges are provided based on structural
properties. Effects of the number and spacing of periods are also investigated.

7.2 Literature review


The issue of what period range to use when computing Saavg has been studied by a
number of researchers. Let Saavg(a, b) denote the value of Saavg computed using a period
range between aT1 and bT1, where a and b are non-negative constants such that a b
and T1 is the fundamental period of the structure. Bianchini (2008) and Bianchini et al.
(2009) studied the displacement response of SDOF and MDOF systems with ductility
levels () ranging from = 1 to 6 using Saavg as the IM. The authors found that the
minimum dispersion in the results was produced using b values of approximately 2 (i.e.,
using twice the value of T1 as the upper limit of the period range) for low ductility levels
and approximately 3 for high ductility levels ( 4). They found that a = 1 was
appropriate for SDOF systems and an a value of approximately 0.25 was best for MDOF
systems with significant contributions from higher modes.

227
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 228

Bianchini (2008) also examined the how the period spacing and number of periods used
to compute Saavg affected the IMs efficiency. He found that using 100 periods to
compute Saavg did not significantly affect the results compared to using 10 periods and
used this as the basis for recommending that 10 periods be used to compute Saavg.
Bianchini also found that using logarithmically spaced periods to compute Saavg generally
resulted in lower dispersions compared to using arithmetically spaced periods given that
10 periods were used. He also found that the period spacing had a slight effect on the
optimum b value.

Tsantaki et al. (2012) studied the collapse of non-degrading bilinear SDOF systems
vulnerable to P- effects. The periods of the SDOFs ranged from T1 = 0.1 s to 5.0 s, and
various hysteretic behaviors and effective post-yield stiffness values were considered.
Instead of characterizing the period range in terms of a and b, the authors looked at the
absolute width of the period range (i.e., the value of T1(b-a)). Using a = 1, the authors
found that the optimum width of the period range (that which produced the minimum
dispersion in the collapse intensities) was

0.2, 0.2 s T1 < 0.5s


0.4, 0.5s T1 < 0.7 s

T1 (b a ) = (7.1)
0.8, 0.7 s T1 < 1.5s
1.2, 1.5s T1 < 5.0 s

The authors found that Saavg did not reduce the dispersion compared to Sa(T1) for systems
with T1 = 0.1 s. In a related study using the same SDOFs but focusing on systems with
bilinear hysteretic behavior, Tsantaki and Adam (2013) recommend using b = 1.6 for
systems with T1 0.15 s and b = 10.67T1 when T1 0.15 s, where a = 1 in all cases.
They found that the optimum b value was not very sensitive to the damping ratio or the
effective post-elastic stiffness.

Bojrquez and Iervolino (2011) also studied the optimum period range for Saavg by using
a vector IM consisting of Sa(T1) and the ratio of Saavg to Sa(T1). Using a = 1, they found
using a b value of approximately 2 to 2.5 provided the lowest dispersion in maximum
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 229

interstory drift ratios based on case studies of a 5-story RC MRF structure subjected to
both ordinary and pulse-like ground motion sets and an 8-story steel MRF structure
subjected to narrow-band ground motions.

7.3 Effect of period spacing and number of periods on the value of Saavg
This section provides an illustrative example of how the period spacing scheme
(specifically arithmetic versus logarithmic) and number of periods used within a given
period range affect the value of Saavg. This example is used to motivate subsequent
sections that examine how the spacing and number of periods affects the dispersion on
the collapse intensities measured by Saavg and the period range that produces the overall
minimum dispersion for a given structure. Following the notation used in the previous
section, let Saavg(a, b) denote the value of Saavg computed using a period range between
aT1 and bT1, where a and b are non-negative constants such that a b and T1 is the
fundamental period of the structure.

To illustrate how the spacing and number of periods used affects the value of Saavg, a
ground motion from the 1999 Chi-Chi, Taiwan earthquake is used as an example. Figure
7.1 shows the pseudo-acceleration spectrum the ground motion. Markers on the spectrum
indicate the periods used to compute Saavg(0.2, 3) for T1 = 1.33 s when either 10 or 100
periods are considered. The periods in Figure 7.1(a) are arithmetically (i.e., uniformly)
spaced within the period range while the periods in Figure 7.1(b) are spaced
logarithmically within the period range. Horizontal lines in Figure 7.1 indicate the value
of Saavg(0.2, 3) computed using the different methods.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 230

Figure 7.1: Effect of period spacing scheme and number of periods used to compute
Saavg(0.2, 3) for T1 = 1.33 s: (a) arithmetic spacing; (b) logarithmic spacing.

As indicated in Figure 7.1, logarithmic spacing samples more heavily from the lower half
of the period range, and the value of Saavg computed with logarithmic spacing tends to be
higher than the value computed with arithmetic spacing because, outside of the short
period range, spectral acceleration values tend to decrease with increasing period. The
relative difference between the values of Saavg computed using arithmetic versus
logarithmic spacing depends on the particular ground motion and period range of interest.
For the ground motion shown in Figure 7.1, computing Saavg using 100 periods spaced
arithmetically versus logarithmically results in Saavg(0.2, 3) = 0.72 g versus 1.24 g,
respectively, for T1 = 1.33 s. This represents the most extreme difference from the ground
motions in the MRCD 137 set, which is documented in Appendix A, as the difference for
this particular period range is approximately 0.10 g, on average, for this ground motion
set. The difference in the value of Saavg computed with 10 or 100 periods for the same
spacing scheme is generally smaller than the difference due to the spacing scheme.
Again, the ground motion in Figure 7.1 represents the most extreme difference in these
values observed for the MRCD 137 ground motion set. The value of Saavg(0.2, 3)
computed using 100 periods versus 10 periods is 0.01 g higher, on average, for
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 231

logarithmic spacing, and there is no difference, on average, for arithmetic spacing for the
MRCD 137 ground motion set. Because of potential differences in the value of Saavg
computed using different spacing schemes and numbers of periods, it is important that the
methods of computing Saavg for the hazard curve and to describe the structural response
conditioned on intensity are consistent. This is in addition to ensuring that the ground
motion orientation used to compute Saavg (e.g., single component versus geometric mean
of two horizontal components) is consistent. Baker and Cornell (2006b) study the latter
case for Sa(T1) and provide examples quantifying the error introduced into risk-based
structural assessments when using inconsistent orientation representations.

7.4 Effect of period range used to compute Saavg on the efficiency


This section examines how the period range used to compute Saavg affects the efficiency
of the IM and the collapse risk results. Primary focus is on the effect of efficiency, and
this is examined using case studies of SDOF and MDOF systems, the latter of which
include generic moment-resisting frame (MRF) and shear wall structures as well as
reinforced concrete (RC) MRFs encompassing a range of periods and post-yield
behavior. The effect of the period range used to compute Saavg on the collapse risk results
is also considered using a case study structure. The impact of period spacing is also
explored here as it is shown to affect the period range that produces the maximum
efficiency. As in previous sections, Saavg(a, b) is computed using periods between aT1
and bT1. The notations a* and b* are used to denote the values of a and b, respectively,
for which Saavg(a, b) produces the minimum dispersion in the collapse intensities for a
given structure. In these studies, values of a between 0 and 1 and values of b between 1
and 10/T1 were considered (i.e., the lower limit of the period range was between 0 and T1,
and the upper limit was between T1 and 10 s). A limit of 10/T1 was placed on the b value
for practical reasons, primarily due to a general lack of ground motion prediction
equations that predict spectral values at periods longer than 10 s that permit the
computation of seismic hazard data needed for collapse (and other structural response-
related) risk assessments.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 232

7.4.1 Case study structure


The case study of the 4-story steel MRF analyzed with the MRCD 137 ground motion set
described in Section 3.5 is used as an illustrative example of how the period range used to
compute Saavg affects the efficiency of the IM. This structure has a fundamental period of
T1 = 1.33 s. Discussions on the efficiency of Saavg(a, b) for the special case where a = b
and of how the period range affects the computed collapse risk are included in the
subsection focusing on arithmetic spacing

7.4.1.1 Arithmetic period spacing


Figure 7.2 shows a contour plot of the dispersion (ln) on the collapse intensities
measured by Saavg(a, b) as a function of a and b when Saavg is computed using
arithmetically spaced periods. The minimum dispersion of ln = 0.22 is achieved with
Saavg(0.2, 3) (i.e., a* = 0.2 and b* = 3). Figure 7.2 shows that the efficiency of Saavg is
generally significantly more sensitive to the upper limit of the period range (the b value)
than the lower limit of the period range (the a value). This figure also demonstrates that a
reduction in dispersion is generally achieved when periods less than T1 (i.e., a < 1) are
used to compute Saavg. For example, the dispersion decreases from ln = 0.29 for
Saavg(1, 3) to ln = 0.22 for Saavg(0.2, 3). The dispersion of the collapse intensities using
Sa(T1) as the IM is ln = 0.39 and can be seen in Figure 7.2 as Sa(T1) is equivalent to
Saavg(1, 1). Figure 7.2 demonstrates that there are a wide range of periods (i.e., of
combinations of a and b) for which Saavg is more efficient than Sa(T1).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 233

Figure 7.2: Contour plot of the dispersion (ln) on the collapse intensities measured by
Saavg(a, b) as a function of the a and b values used to compute Saavg using arithmetic
period spacing.

The contours in Figure 7.2 suggest that considering Saavg(a, b) for a > 1 when a < b will
not increase the efficiency of the IM. Here the special case of Saavg(a, b) where a = b is
considered, which is equivalent to considering Sa(aT1). The advantages of this special
case (i.e., spectral acceleration at a single period) are that the earthquake engineering
community is familiar with this type of IM and that seismic hazard data, including
probabilistic seismic hazard analysis (PSHA) deaggregations, is generally available (e.g.,
USGS 2012a). Figure 7.3 shows the dispersion of collapse intensities measured by
Saavg(a, b) when a = b as a function of the a value for the case study from Section 3.5.
The minimum dispersion of ln = 0.29 is achieved when a = 1.35, which represents an
approximately 25% reduction in dispersion compared to using Sa(T1) (i.e., a = 1). A
reduction in the dispersion of collapse intensities measured by Sa(aT1) for a > 1
(compared to a = 1) is consistent with the findings of Haselton and Baker (2006), who
studied collapse using a set of 1-s SDOF systems. The authors defined collapse as the
exceedance of a ductility limit and found that the a values associated with the minimum
dispersion ranged from 1.0 to 2.9. Based on the results from all systems, the authors
recommended using an a value of approximately 2.

Vamvatsikos and Cornell (2005a, Chapter 6) also investigated using Sa(aT1) to reduce
the dispersion of collapse intensities compared to that calculated using Sa(T1). They
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 234

defined collapse as numerical non-convergence. The authors found that a > 1 did not
reduce the dispersion in collapse intensities compared to using Sa(T1) for either a 9-story
or a 20-story steel MRF structure. For the 9-story structure, which had a fundamental
period of T1 = 2.4 s, the authors found that periods other than T1 (i.e., a values other than
1) did not result in a significant reduction in dispersion compared to that found using
Sa(T1). For the 20-story structure, which had a fundamental period of T1 = 4.0 s, the
authors found that the minimum dispersion in collapse intensities occurred at an a value
of approximately 0.6, which reduced the dispersion by approximately 30% compared to
using Sa(T1). For a 5-story steel braced-frame structure, however, which had a
fundamental period of T1 = 1.8 s, the authors found that the minimum dispersion in
collapse intensities occurred at an a value of approximately 1.2, which reduced the
dispersion by approximately 30% compared to using Sa(T1). Vamvatsikos and Cornell
concluded that using an a value greater than 1 could reduce the dispersion in collapse
intensities for structures with insignificant contributions from higher modes but that for
other types of structures the values of a that resulted in decreased dispersion could be
difficult to identify a priori.

Figure 7.3: Dispersion of collapse intensities measured by Saavg(a, b) using arithmetic


period spacing for the special case of a = b, which is equivalent to Sa(aT1), as a function
of a.

Though Figure 7.3 shows that the reduction in dispersion using Sa(aT1) for a > 1
(compared to a = 1) can be significant, Figure 7.2 shows that there are many periods
ranges for which Saavg(a, b) produces smaller dispersions than ln = 0.29, the minimum
dispersion found for Sa(aT1). Though not shown, evaluation of using a single
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 235

elongated period as the IM was repeated using the six other ground motion sets from
Section 6.4, and it was found that Sa(aT1) could generally reduce the dispersion in
collapse intensities by approximately 20% compared to Sa(T1) and that the a value
associated with the minimum dispersion was typically on the order of 1.3. Given that
Saavg(a, b) for a < b can typically reduce the dispersion more than a = b, the remainder of
this chapter focuses on the former case.

The dispersion of the collapse intensities is not the only quantity relevant to the collapse
risk that is affected by the period range used to compute Saavg: the median collapse
intensity and the seismic hazard curve also vary as a function of the period range.
Changes in the seismic hazard curve are caused by changes in both the expected value
and the dispersion of Saavg predicted by the ground motion prediction equation (GMPE).

To investigate how sensitive the collapse risk estimates are to the period range used to
compute Saavg, the collapse risk assessment case study described in Section 6.4 is
repeated using the MRCD 137 ground motion set. Figure 7.4 shows contours of the
collapse risk estimates given by Saavg(a, b) as a function of the a and b values used to
compute Saavg. The mean annual frequency of collapse (c) in units of 10-4 is shown in
Figure 7.4(a), and the probability of collapse in 50 years (Pc,50) expressed a as percent is
shown in Figure 7.4(b).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 236

Figure 7.4: Contours of collapse risk estimates measured by Saavg(a, b) as a function of


the a and b values used to compute Saavg: (a) mean annual frequency of collapse, c, in
units of 10-4; (b) probability of collapse in 50 years, Pc,50, expressed as a percent.

As seen in Figure 7.4 the collapse risk estimates do vary as a function of the period range
used to compute Saavg. For example, Figure 7.4(b) shows that Pc,50 is approximately 0.3%
for Saavg(0.2, 3) and approximately 0.6% for Saavg(1, 2). Reasons for the increased
collapse risk estimate given by Saavg(1, 2) include the slight increase in the dispersion of
the collapse intensities for this period range (lnSaAvg(0.2,3) = 0.22 versus lnSaAvg(1,2) = 0.25
as seen in Figure 7.2), although it is believed a more significant reason is the increased
dispersion of the Saavg value predicted by the GMPE (lnSaAvg,GMPE). Using the 2008 Boore
Atkinson (BA08) GMPE (Boore and Atkinson 2008) and considering an unspecified fault
type, the value of lnSaAvg,BA08 is 0.59 for Saavg(0.2, 3) versus 0.67 for Saavg(1, 2). As
discussed in Section 6.4, the dispersion on the predicted value of ln(Sa(T)) (lnSa(T),GMPE)
generally increases with period for the BA08 GMPE, which was used to compute the
seismic hazard curve. These dispersions vary from a minimum of lnSa(T),BA08 = 0.57 at T =
0 s to a maximum of lnSa(T),BA08 = 0.78 at T = 7.5 s. The values of lnSa(T),GMPE are used to
compute lnSaAvg,GMPE as shown in Equation 6.10 and therefore explain differences in the
value of lnSaAvg,GMPE as a function of the period range used. The expected values of Saavg
predicted by the GMPE that are used to calculate the seismic hazard curve also change as
a function of the period range used to compute Saavg, and this can also affect the collapse
risk estimates.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 237

It is very important to note that the sufficiency of Saavg(a, b) with respect to ground
motion parameters such as magnitude, distance, and spectral shape was not considered
for the different period ranges shown in Figure 7.4. This is noted as the sufficiency of
Saavg(a, b) can also affect the collapse risk estimate, potentially more significantly than
changes in the dispersion of the collapse intensities and the mean and standard deviation
values predicted by the GMPE as a function of the period range. As previously discussed
in Chapters 5 and 6, the ground motions used to analyze the structure can produce biased
response estimates when the IM is insufficient with respect to one or more ground motion
parameters. Section 6.3 demonstrated that Saavg(0.2, 3) was generally sufficient with
respect to magnitude, distance, and spectral shape measures, and this was reinforced by
the findings of Section 6.4 that showed the collapse risk estimate measured by
Saavg(0.2, 3) was relatively insensitive to the ground motion set used to analyze the
structure. However, Section 5.6 showed that the collapse intensities measured by
Saavg(1, 1), which is equivalent to Sa(T1), generally depend on the spectral shape of the
ground motions, and this was also reflected in the findings of Section 6.4 that showed the
collapse risk estimate measured by Sa(T1) could vary significantly depending on the
ground motion set used. For example, when the collapse risk estimate measured by
Sa(T1) was adjusted to be consistent with the expected spectral shapes of records at the
intensity with the greatest contribution to the collapse risk as described in Section 6.4 (or
when directly using records with the expected distribution of spectral values at this
intensity), the collapse risk estimate dropped significantly from Pc,50 1.0% to Pc,50
0.4% (see Figure 6.16(c)), which is much closer to the collapse risk predicted by
Saavg(0.2, 3) of Pc,50 0.3% as shown in Figure 7.4(b) (and also in Figure 6.16(c)). In
other words, this suggests that after correcting for any bias in the structural response
estimates introduced by using an insufficient IM with ground motions that are not
representative of the motions expected at the site (i.e., not consistent with the seismic
hazard), the collapse risk estimates measured by Saavg(a, b) shown in Figure 7.4 may not
exhibit as much variability with the period range used to compute Saavg as is indicated in
the figure.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 238

This brings up the very important topic of hazard consistency in estimating the risk
associated with structural responses such as collapse. It has been recently argued (NIST
2011, Appendix A; Bradley 2012; Lin 2012, Chapter 4; Lin et al. 2013b) that the seismic
risk of a given structure at a given site is unique and theoretically not dependent on the
IM one uses in risk assessment. The sources previously cited state that when ground
motions are carefully selected to be consistent with the seismic hazard (i.e., the ground
motions are representative of the distribution of ground motion parameters affecting
structural response that are expected at a given intensity level), the computed risk should
be the same no matter what IM is used. Though the ground motions used in Section 6.4
and again this section were not selected using the rigorous methodologies discussed in
the previous sources, which are described in subsequent paragraphs (e.g., the records
were not reselected at different intensity levels to reflect changes in the earthquake
magnitudes dominating the seismic hazard), the collapse risk estimates for the case study
measured by Saavg(0.2, 3) and by Saavg(1, 1) (or, equivalently, Sa(T1)) with the spectral
shape adjustment presented in Figure 6.16(c) are generally in relatively good agreement,
although Saavg(0.2, 3) clearly produces more stable collapse risk estimates among the
different ground motion sets considered.

Examples presented in Appendix A of NIST Report GCR 11-917-15 (2011), which


appear in more detail in Lin (2012, Chapter 4 and Appendix A) and Lin et al. (2013b),
select ground motions at a given intensity level based on their match to a conditional
spectrum (CS) that describes the distribution (both mean and logarithmic standard
deviation) of spectral values at all periods based on a PSHA deaggregation. It is shown
that when different conditioning periods are used to derive the CS and characterize the
seismic hazard (e.g., using Sa(T1) versus Sa(2T1) as the IM), the resulting risk estimates
(i.e., mean annual frequency of exceeding a peak drift or peak floor acceleration, or c)
derived from different conditioning periods are typically in good agreement. For the
example 20-story RC MRF structure with T1 = 2.6 s used in Lin (2012, Chapter 4) and
Lin et al. (2013b), the c estimates vary from c = 3.12 x 10-4 for a conditioning period of
T = 0.45 s (which is equal to T3) to c = 5.02 x 10-4 for a conditioning period of T1 = 2.6 s
(Pc,50 from approximately 1.5% to 2.5%, respectively), which the authors consider a
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 239

small difference for the range of conditioning periods investigated. The authors also note
that approximations are involved in constructing the CS (as an exact CS that accounts for
multiple GMPEs used to construct the seismic hazard curve and multiple seismic sources
that contribute to the hazard, which implies additional variability of spectral values, was
not practical to compute at the time) but maintain that by selecting ground motions that
are truly consistent with the exact CS and other non-spectral ground motion parameters
of interest one should obtain a consistent (and accurate) risk prediction using any
conditioning period. Lin and Lin et al. note that there is still a benefit to choosing a
conditioning period that results in good structural response predictions as it reduces the
number of required analyses for a given level of accuracy (efficiency) and reduces the
impact of any inaccuracies representing spectral values other than T1 (sufficiency).

Bradley (2012) approaches hazard consistency in ground motion selection and risk
assessment via a six step process. A key component of this process is that the ground
motions selected at a given intensity level must match the distribution of ground motion
parameters affecting structural response that is expected based on the value of the
conditioning IM used to characterize the seismic hazard. This requires knowledge of
GMPEs and correlations between the different ground motion parameters used to form
the multivariate distribution of ground motion parameters and is achieved by use of the
generalized conditional intensity measure (GCIM) approach, which is described in detail
by Bradley (2010). Applying the six step process to a case study of a bridge-foundation-
soil system and considering 14 different ground motion parameters (including spectral
ordinates, peak ground velocity and acceleration, and duration measures), Bradley (2012)
showed that the seismic demand hazards of peak pile displacement, peak pile curvature,
and peak deck acceleration were in relatively good agreement using six different
conditioning IMs, including PGA, PGV, and spectral ordinates. Furthermore, Bradley
found that at the 2%-in-50-year hazard level neither the peak pile head displacement nor
the peak deck acceleration hazards estimated by the different conditioning IMs could be
distinguished as statistically different at the 5% significance level. Bradley notes that of
the six main sources of error he considers in the numerical computation of the seismic
demand hazard in a practical context, one of the major sources of uncertainty in the
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 240

computed hazard is related to using a finite number of ground motions. He also notes that
there is relationship between uncertainty in the seismic demand hazard and the
uncertainty in the seismic response conditioned on intensity measure (i.e., the lognormal
standard deviation) and recommends that the conditioning IM be chosen so that it
captures the amplitude of ground motion in the region of vibration frequencies that
govern the predominant physical mechanism controlling the seismic response of the
problem considered.

The findings of Bradley (2012), Lin (2012, Chapter 4) and Lin et al. (2013b) support the
conclusion that though ensuring consistency between the seismic hazard and ground
motions used in response analysis may theoretically lead to the same risk prediction
independent of the conditioning IM, in a practical context there is still value in using an
efficient IM as reduced uncertainty in the structural response should lead to a more
accurate prediction of the collapse risk. Keeping this in mind, the remainder of this
chapter focuses on identifying how changes in the period range affect the efficiency of
Saavg and investigating trends in the period ranges that produce the minimum dispersion
in collapse intensities (i.e., the maximum efficiency) for different structures, which is also
useful when structural response prediction is of primary concern. Additionally, although
it is theoretically possible to find a set of ground motions that match the probability
distribution of the conditioning IM, in reality one may encounter problems finding
recorded ground motion that fit the target probability distribution. In such cases, using an
IM that is robust with respect to changes in ground motion set is a very desirable
characteristic.

Before considering other structures, the case study structure described in Section 3.5 is
used to illustrate how different ground motion sets can affect the efficiency of Saavg(a, b)
and the optimum period range. Contours of dispersion on the collapse intensities
measured by Saavg(a, b) as a function of the a and b values used to compute Saavg are
presented in Figure 7.5 for different ground motion sets. These ground motion sets were
also used in Section 6.4, and relevant details about the sets are provided there. As
illustrated in Figure 7.5 the minimum dispersion of Saavg ranges from lnSaAvg = 0.20 at
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 241

Saavg(0.6, 2.5) for the Set #1A FN/FP ground motions (Figure 7.5(c)) to lnSaAvg = 0.29 at
Saavg(0.4, 2.8) for the CS, (T1) = 1.7 ground motions (Figure 7.5(e)). The value of a*
varies from 0 to 0.6 between the ground motion sets, with an average value of 0.3;
however, the value of b* is close to 3 (between 2.5 and 3.2) for all ground motion sets
considered (Figure 7.2 and Figure 7.5). It is important to highlight that even though no
single period range produces the minimum dispersion for all ground motion sets, the
dispersion using Saavg(0.2, 3) is no more than 0.03 lnSaAvg units greater than the minimum
dispersion for each ground motion set.

Figure 7.5: Contours of dispersion (ln) on the collapse intensities measured by Saavg(a, b)
as a function of the a and b values used to compute Saavg using arithmetic period spacing
for different ground motion sets: (a) LMSR-N; (b) Set One; (c) Set #1A, FN/FP; (d) Set
#1A, unrotated; (e) CS, (T1) = 1.7; (f) CS, (T1) = 1.8.

7.4.1.2 Logarithmic period spacing


The discussion now turns to how the number and spacing of periods affects the dispersion
of the collapse intensities measured by Saavg(a, b). Figure 7.6 shows contours of the
dispersion (ln) on the collapse intensities measured by Saavg(a, b) as a function of the a
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 242

and b values used to compute Saavg. Figure 7.6(a) and Figure 7.6(b) present the results
when Saavg is computed using 100 periods and 10 periods, respectively, spaced
arithmetically between aT1 and bT1 1, which shows that in general neither the dispersion
associated with a particular Saavg(a, b) nor the values of a* and b* associated with the
minimum dispersion are sensitive to how many periods are used. A similar result is
observed when using logarithmic spacing, as the contours shown in Figure 7.6(c) based
on 100 periods are not significantly different from those in Figure 7.6(d), which are based
on 10 periods. Studies of other structures show that this observation is generally true for
both arithmetic and logarithmic spacing. Note that although there is a shift in the a* and
b* values from Figure 7.6(c) to Figure 7.6(d), which are denoted by a star in these
figures, the contour based on ln = 0.23 does not shift significantly, indicating that either
set of a* and b* values would produce essentially the same dispersion.

1
Note that in previous chapters Saavg was previously computed as Saavg(0.2, 3) using periods arithmetically
spaced every 0.01 s, which happens to involve 373 periods for this structure. The results are the same
whether 100 or 373 arithmetically spaced periods are used to compute Saavg(0.2, 3).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 243

Figure 7.6: Contours of the dispersion (ln) of the collapse intensities as a function of the
a and b values used to compute Saavg(a, b) for various period spacing schemes:
(a) arithmetic spacing with 100 periods; (b) arithmetic spacing with 10 periods;
(c) logarithmic spacing with 100 periods; (d) logarithmic spacing with 10 periods.

Both the arithmetic and logarithmic period spacing schemes achieve an overall minimum
dispersion of lnSaAvg(a*,b*) = 0.22 as shown in Figure 7.6; however, the a* and b* values
associated with the minimum dispersion for a given period spacing are not the same.
Whereas Saavg(0.2, 3.0) produces the minimum dispersion for arithmetic spacing when
using 100 periods, Saavg(0.5, 3.7) produces the minimum dispersion for logarithmic
spacing when using 100 periods. The increase in both the a* and b* values for
logarithmic spacing relative to arithmetic spacing is also seen when looking at the
dispersion contours for this structure using the six other ground motion sets from Section
6.4, which are presented in Figure 7.7. Compared to similar contours presented in Figure
7.5 for arithmetic spacing, it is seen that the values of a* and b* increase by an average of
approximately 50% and 20%, respectively, for logarithmic spacing. The increase in a*
for logarithmic relative to arithmetic spacing is because logarithmic spacing samples
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 244

more heavily from the left half of the period range, and the spectral values at periods near
or greater than the fundamental period of the structure tend to capture more of the
damage potential (i.e., damaging pulses) of the ground motion than values at periods that
are short relative to the fundamental period. The increase in b* values may be related to
the increasing a* values, as for a given value of a increasing the b value will reduce the
concentration of periods near the a value for logarithmic spacing.

Figure 7.7: Contours of dispersion (ln) on the collapse intensities measured by Saavg(a, b)
as a function of the a and b values used to compute Saavg with logarithmic period spacing
for different ground motion sets: (a) LMSR-N; (b) Set One; (c) Set #1A, FN/FP; (d) Set
#1A, unrotated; (e) CS, (T1) = 1.7; (f) CS, (T1) = 1.8.

7.4.2 Generic SDOF systems


This section examines the optimum period range for a generic set of SDOF systems when
using arithmetic period spacing. The intent of this study is to determine if the trends of
reduced dispersion in collapse intensities using Saavg(a, b) versus Sa(T1) that are observed
for MDOF systems are also observed for SDOF systems. Particular focus is given to what
benefit, if any, is gained by including periods less than the fundamental period (i.e., using
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 245

a < 1) as previous studies (e.g., Bianchini 2008; Tsantaki et al. 2012) have generally not
discussed this for SDOF systems.

Periods of 0.1 s, 0.2 s, 0.5 s, 0.75 s, 1.0 s, 1.5 s, 2.0 s, 2.5 s, 3.0 s, 3.5 s, and 4 s are
considered, and 5% damping is used for all systems. A bilinear hysteretic model is used
and the force-deformation behavior is governed by a tri-linear backbone curve as
illustrated in Figure 7.8. All systems have a yield base shear coefficient of Vy/W = 0.25
(the yield base shear normalized by the weight of the system), a 3% strain hardening ratio
(i.e., the post-yield stiffness is 3% of the elastic stiffness), a plastic deformation of p =
4y as seen in Figure 7.8 (where y is the yield deformation), a post-capping deformation
of pc = 7y as seen in Figure 7.8, and a stability coefficient of = 0.03 (the weight of the
system times the lateral displacement normalized by the product of the height of the
system and base shear at which the lateral displacement is measured). The bilinear
hysteretic model uses the modified Ibarra-Medina-Krawinkler deterioration model
(Lignos and Krawinkler 2012a, Chapter 2) and all cyclic deterioration parameters are
defined as = 50. Incremental dynamic analyses are performed in IIIDAP (Lignos 2009)
using the LMSR-N ground motion set.

Figure 7.8: Tri-linear backbone curve used to define SDOF systems.

Contour plots showing the dispersion of collapse intensities measured by Saavg(a, b)


computed using arithmetic period spacing are presented for the different SDOF systems
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 246

in Figure 7.9. As shown in this figure, the dispersion is typically much more strongly
influenced by the b value or the upper limit of the period range, which was also observed
for MDOF case study structure. Most systems presented in Figure 7.9 achieve the
minimum dispersion at b values between 2 and 3. Note that again an upper limit of 10 s
was imposed on the period range used to compute Saavg, which is why some of the
contours of the longer period systems to not extend all the way to b = 6. Based on the
shape of the contours, however, it seems that the value of b* would likely not increase if
periods beyond 10 s were included in the period range. The more significant finding
regarding the period range, however, is that all of the 11 SDOF systems presented in
Figure 7.9 show that there is a benefit to including periods less than the fundamental
period as the minimum dispersion is achieved at a values less than 1 in all cases. This
may be a counterintuitive result for those who subscribe to the theory that accounting for
spectral ordinates at periods shorter than the fundamental is necessary to account for
higher mode effects and, therefore, for an SDOF system there would be no need or
benefit to including periods to the left of T1 when computing Saavg; however, as clearly
shown by these results, including information about spectral ordinates at periods smaller
than the fundamental period leads to reductions in dispersions even for SDOF systems
(which do not have higher modes of vibration).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 247

Figure 7.9: Contours of dispersion (ln) on the collapse intensities measured by Saavg(a, b)
as a function of the a and b values used to compute Saavg for various SDOF systems:
(a) T = 0.1 s; (b) T = 0.2 s; (c) T = 0.5 s; (d) T = 0.75 s; (e) T = 1.0 s; (f) T = 1.5 s;
(g) T = 2.0 s; (h) T = 2.5 s; (i) T = 3.0 s; (j) T = 3.5 s; (k) T = 4.0 s.

As discussed in Section 5.6, it is believed that the reduction in dispersion is not


exclusively due to including spectral ordinates with periods close to those of higher
modes but also, and perhaps more importantly, because including spectral ordinates at
periods less the fundamental period in the spectral averaging range provides information
of critical acceleration pulses in the ground motion that increase displacement demands in
the system. As discussed in Chapter 5, using an average spectral value also minimizes the
role that the initial conditions play in the response of the elastic system, which is
important as a nonlinear system will, in general, not have the same initial conditions as
the elastic system and can therefore respond very differently than the elastic system to the
same ground motion segment. For this reason, it makes sense to take the average spectral
value using periods smaller and larger than the fundamental period as neighboring
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 248

systems (i.e., systems with periods similar to the fundamental period) can represent how
the elastic system may have responded given different initial conditions even for SDOF
systems (that, by definition, do not have higher modes of vibration). Additional details
regarding this reasoning are provided in Sections 5.4 and 5.5.

Further studies using SDOF systems are necessary to determine how the a value
associated with the minimum dispersion of collapse intensities of Saavg(a, b) varies as a
function of system parameters (e.g., less ductile systems than the ones examined here)
and period, but the results presented in Figure 7.9 suggest that even for SDOF systems a
reduction in the dispersion of collapse intensities measured by Saavg is achieved by
including periods less than the fundamental period in the averaging range.

7.4.3 Generic moment-resisting frame structures


Nearly 400 generic MRF structures used by Zareian and Krawinkler (2009) are studied
here to develop recommendations for the range of periods to use when computing Saavg
and to investigate the effect of period spacing when computing Saavg. Structures of 4, 8,
12 and 16 stories are included, and the fundamental period of these structures for a given
number of stories (N) was set to be either T1 = 0.1N, 0.15N or 0.2N seconds. Structural
parameters such as the ductility-dependent strength reduction factor (R), the variation of
strength and stiffness over the height, and the post-elastic behavior of components are
also varied. All structures were analyzed with the LMSR-N ground motion set.
Additional details about these structures are provided in Appendix B.

7.4.3.1 Arithmetic period spacing


Figure 7.10 presents contours of the dispersion of the collapse intensities as a function of
the a and b values used to compute Saavg(a, b) that are representative of results for the 4-
story structures (Figure 7.10(a)), the 8-story structures (Figure 7.10(b)), the 12-story
structures (Figure 7.10(c)), and the 16-story structures (Figure 7.10(d)). Let a* and b*
denote the values of a and b for which Saavg(a, b) produces the minimum dispersion in
the collapse intensities. As indicated by Figure 7.10 a* is 0 or practically 0 for these
structures, which indicates the minimum dispersion is achieved when the period range
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 249

starts from T = 0. In fact, all but 21 of the 396 structures have a* 0.05, and these 21
structures for which a* > 0.05 are all 4-story structures. Figure 7.10 also indicates that the
value of b* generally decreases as the number of stories increases. This trend was
observed although notable exceptions to this trend are the 4-story structures with
T1 = 0.4 s, for which the mean value of b* is 2.6.

Figure 7.10: Contours of the dispersion (ln) of the collapse intensities as a function of the
a and b values used to compute Saavg(a, b) using arithmetic period spacing that are
representative of results for generic MRF structures: (a) 4 story; (b) 8 story; (c) 12 story;
(d) 16 story.

The trends concerning the a* and b* values implied by Figure 7.10 are better illustrated
in Figure 7.11, which shows histograms representing the joint distribution of the a* and
b* values for the 396 generic MRF structures. The histograms are separated based on
number of stories, and for a given number of stories there are 99 structures. As previously
noted, the 4-story structures shown in Figure 7.11(a) tend to have more scatter in the a*
values than the other structures, which all have a* 0.05; however, the majority of a*
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 250

values for the 4-story structures are still less than or equal to 0.05, indicating that cases in
which a* > 0.05 are not only a small fraction of the total number of buildings analyzed
but are even a small fraction of the 4-story buildings. The histograms in Figure 7.11 also
show that for the four building heights the b* values tend to be concentrated between 2
and 4.

Figure 7.11: Histogram of a* and b* values associated with the minimum dispersion of
Saavg(a, b) using arithmetic period spacing for generic MRF structures: (a) 4 story; (b) 8
story; (c) 12 story; (d) 16 story.

Figure 7.12 shows the b* values for each of the 396 structures. The horizontal axis in
Figure 7.12 that displays the structure number (a value between 1 and 396) was added to
make the data easier to see because the points are tightly clustered in some regions,
especially for b* between 3 and 4, which indicates that these period ranges generally give
the minimum dispersion for the generic MRF structures examined here. The vertical lines
divide the structures into bins based on fundamental period. For a given number of
stories, the left, center, and right bins indicate structures with fundamental periods of T1 =
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 251

0.1N, 0.15N or 0.2N seconds, respectively. Figure 7.12 shows that with the exception
of the 4-story structures with T1 = 0.1N (i.e., Structure 1 through Structure 33), the value
of b* for a given number of stories tends to decrease as the fundamental period of the
structure increases, and this trend generally becomes more pronounced as the number of
stories increases.

Figure 7.12: Scatter plot of b* using arithmetic period spacing for all generic MRF
structures.

To investigate if there are any trends with b* as a function of T1, Figure 7.13 shows the
b* values of all 396 structures plotted as a function of T1 in the log-log domain. The
correlation coefficient between b* and T1 is = -0.58, indicating that b* generally
decreases with increasing T1. This correlation coefficient was computed for all structures
excluding those with T1 = 0.4 s, which as noted earlier do not follow the trend of
decreasing b* with increasing T1. A linear regression to the data shown Figure 7.13
indicates that the b* value is approximately 4.1 for structures with T1 = 0.6 s and
decreases to approximately 2.5 when T1 = 3.2 s. As seen in Figure 7.13 there is a
considerable amount of scatter about the regression, which is reflected by the correlation
coefficient of = -0.58 (or equivalently the 2 = R2 value of 0.34).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 252

Figure 7.13: b* value associated with the minimum dispersion of Saavg(a, b) as a function
of T1 using arithmetic period spacing for the generic MRF structures.

Other trends between the value of b* and structural properties (e.g., strength, stiffness
distribution, and post-yield structural component parameters) are examined. Though
some minor trends are observed (e.g., for systems whose components have small plastic
deformation capacities, the value of b* generally increases with increasing values of R,
with all other structural properties being equal), the relationship between b* and the
structural properties appears to be primarily dominated by the fundamental period of the
system.

To develop a general and simple recommendation for the values of a and b to use when
computing Saavg that are not a function of structural properties such as T1, the difference
between the dispersion on the collapse intensities computed using Saavg(a*, b*) (i.e., the
overall minimum dispersion for each structure) versus Saavg(a, b) for given values of a
and b were computed for each of the 396 generic MRF structures. The mean value of this
difference for all 396 structures is shown in Figure 7.14, which shows a contour plot of
the difference in dispersion as a function of the a and b values used to compute
Saavg(a, b). Figure 7.14(a) presents the results when the difference is computed in
absolute terms (e.g., lnSaAvg(a,b) - lnSaAvg(a*,b*)), and Figure 7.14(b) presents the results
when the absolute difference is normalized by the value of lnSaAvg(a*,b*). The latter case
essentially represents a percent difference and is used because the overall minimum
dispersion varies between structures from lnSaAvg(a*,b*) = 0.11 for one of the 4-story
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 253

structures to lnSaAvg(a*,b*) = 0.34 for one of the 16-story structures. Both plots in Figure
7.14 indicate that a = 0 and b = 3 are convenient values to use when computing
Saavg(a, b) as these values show a relatively small difference, on average, from the
absolute minimum dispersion that can be calculated for a given structure using Saavg. The
absolute difference is on the order of 0.02 lnSaAvg units when using a = 0 and b = 3 as
seen in Figure 7.14(a), and the relative difference is on the order of 0.1, or 10%, as seen
in Figure 7.14(b).

Figure 7.14: Contour plot showing the mean value, based on the 396 generic MRF
structures, of the difference between the dispersion on the collapse intensities computed
using Saavg(a*, b*) versus Saavg(a, b) as a function of a and b using arithmetic period
spacing: (a) difference in absolute terms; (b) difference normalized by the dispersion
based on Saavg(a*, b*).

Figure 7.15 is used to investigate how the efficiency (as measured by the dispersion of
the collapse intensities) of Saavg(0, b) is affected when b is calculated using the regression
with T1 shown in Figure 7.13 (denoted as b:reg) versus the constant (i.e., independent of
T1) recommendation of b = 3 and how these compare to the efficiency of Saavg(a*, b*).
Figure 7.15(a) is a scatter plot comparing the value of lnSaAvg(a*,b*) to lnSaAvg(0,b:reg) for
each of the 396 generic MRF structures. Note that the b:reg value predicted by the
equation shown in Figure 7.13 is used for structures with T1 = 0.4 s, even though these
structures were not used in determining the regression coefficients. Many points in Figure
7.15(a) lie at or on the x = y line, which indicates that the regression is generally
effective at predicting a b value that, when used with a = 0, will give a dispersion that is
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 254

essentially the same as the overall minimum dispersion for that structure. This is
reinforced by that fact that the point denoting the mean dispersion for all structures is
close to the x = y line. Similar observations are made when comparing lnSaAvg(a*,b*) to
lnSaAvg(0,3), which is shown in Figure 7.15(b). Figure 7.15(c) compares lnSaAvg(0,3) to
lnSaAvg(0,b:reg), and it is seen that computing b using the regression versus a period-
independent value of b = 3 generally leads to a smaller dispersion as most points,
including the mean point, lie above the x = y line. However, because (1) the difference
in the dispersion when b is computed from the regression versus using the period-
independent value of b = 3 is generally relatively small; and (2) using b:reg versus b = 3
actually increases the dispersion for some structures, this indicates that there is generally
not a significant advantage to using b:reg (a period-dependent b) over b = 3.

Figure 7.15: Scatter plot of dispersion in collapse intensities computed using Saavg with
arithmetic period spacing for the 396 generic MRF structures for different computations
of Saavg(a, b): (a) Saavg(a*, b*) vs. Saavg(0, b:reg); (b) Saavg(a*, b*) vs. Saavg(0, 3); (c)
Saavg(0, 3) vs. Saavg(0, b:reg). Note that b:reg is the b value obtained from the regression
based on T1 shown in Figure 7.13.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 255

7.4.3.2 Logarithmic period spacing


Figure 7.16 presents contours of the dispersion of the collapse intensities as a function of
the a and b values used to compute Saavg(a, b) using logarithmic period spacing that are
representative of results for the 4-story structures (Figure 7.16(a)), the 8-story structures
(Figure 7.16(b)), the 12-story structures (Figure 7.16(c)), and the 16-story structures
(Figure 7.16(d)). As indicated by Figure 7.16, a* is generally between 0.2 and 0.4 and the
value of b* generally decreases as the number of stories increases. Periods greater than
10 s are not included in the range to compute Saavg, which is why some of the dispersion
contours in Figure 7.16 do not extended to the maximum b value shown in the figure.
This decision is made as many GMPEs do not predict spectral values at periods greater
than 10 s, which means that even though using a period range extending past 10 s may
reduce the dispersion in the collapse intensities, seismic hazard data, which is necessary
to compute the collapse risk, may not be available or may not be reliable as spectral
ordinates beyond 10 s are typically strongly affected by base-line correction and other
record processing modifications applied to ground motion recordings. The shape of the
dispersion contours suggests that for logarithmic spacing there may be more of a
penalty on the dispersion when underestimating either a* or b* than when
overestimating either a* or b* (i.e., it may be better to overestimate a* or b* than to
underestimate these values because the dispersion appears more sensitive to the latter
case).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 256

Figure 7.16: Contours of the dispersion (ln) of the collapse intensities as a function of the
a and b values used to compute Saavg(a, b) for logarithmic period spacing that are
representative of results for generic MRF structures: (a) 4 story; (b) 8 story; (c) 12 story;
(d) 16 story. Note the different vertical axes between (a) and (b) versus (c) and (d).

To understand in a more general way how the a* and b* values vary for logarithmic
spacing, histograms describing the distribution of a* and b* for the generic MRF
structures are shown in Figure 7.17. Similar to the arithmetic spacing presented in the
previous section, the histograms are broken down by number of stories, where there are
99 structures for a given number of stories. A comparison of the a* and b* histograms for
logarithmic spacing (Figure 7.17) and arithmetic spacing (Figure 7.11) shows that both
the a* and b* values vary much more between structures for logarithmic spacing than for
arithmetic spacing.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 257

Figure 7.17: Histogram of a* and b* values associated with the minimum dispersion of
Saavg(a, b) using logarithmic period spacing for generic MRF structures: (a) 4 story;
(b) 8 story; (c) 12 story; (d) 16 story. Note this figure uses logarithmic period spacing
while Figure 7.11 uses arithmetic period spacing.

Scatter plots of the a* and b* values found using logarithmic spacing versus the
fundamental structural period are shown in the log-log domain in Figure 7.18(a) and
Figure 7.18(b), respectively, for the generic MRF structures. It is seen that there is only a
mild downward trend between a* and T1 as reflected by the correlation coefficient of
= -0.20; however, a significantly stronger linear trend exists between b* and T1 in the
log-log domain as the correlation coefficient is = -0.74. The 4-story structures with T1 =
0.4 s were excluded from both regressions as they generally did not follow the trend,
which was also observed using arithmetic spacing.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 258

Figure 7.18: Relationship between the a* and b* values associated with the minimum
dispersion of Saavg(a, b) obtained using logarithmic period spacing and T1 for the generic
MRF structures: (a) a* and T1; (b) b* and T1.

Figure 7.19 is used to investigate how the efficiency (as measured by the dispersion of
the collapse intensities) of Saavg(a, b) is affected when a and b are calculated using the
regression with T1 shown in Figure 7.18 (denoted as a:reg and b:reg, respectively) versus
the period-independent recommendation of a = 0.2 and b = 9 and how these compare to
the efficiency of Saavg(a*, b*). The values of a and b used for the period-independent
recommendation are the values that, on average based on the 396 structures, result in the
smallest difference between the overall minimum dispersion of a given structure and the
dispersion calculated using period-independent values of a and b. Note that a maximum
period of T = 10 s was used to compute Saavg even if bT1 exceeded 10 s (i.e., b for the
period-independent recommendation was taken as the minimum of 9 and 10/T1). This
was done because many GMPEs do not predict spectral values at periods greater than 10
s, meaning that even if including periods greater than 10 s reduces the dispersion, the
collapse risk may not be able to be computed using Saavg over such a period range
because of a lack of seismic hazard data. Figure 7.19(a) is a scatter plot comparing the
value of lnSaAvg(a*,b*) to lnSaAvg(a:reg,b:reg) while Figure 7.19(b) compares lnSaAvg(a*,b*) to
lnSaAvg(0.2,9) and Figure 7.19(c) compares lnSaAvg(0.2,9) to lnSaAvg(a:reg,b:reg). It can be seen
that both the period-dependent and period-independent values of a and b generally
produce a dispersion close to the overall minimum dispersion of a given structure. Figure
7.19(c) shows that there is generally minimal, if any, an advantage to using the period-
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 259

dependent values of a and b instead of the period-independent values of a = 0.2 and


b = 9.

Figure 7.19: Scatter plot of dispersion in collapse intensities computed using Saavg with
logarithmic period spacing for the 396 generic MRF structures for different computations
of Saavg(a, b): (a) Saavg(a*, b*) vs. Saavg(a:reg, b:reg); (b) Saavg(a*, b*) vs. Saavg(0.2, 9);
(c) Saavg(0.2, 9) vs. Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a and b values,
respectively, obtained from the regression based on T1 shown in Figure 7.18.

7.4.3.3 Comparison of period spacing schemes


Figure 7.20 plots the values of lnSaAvg(a*,b*) computed using arithmetic versus logarithmic
spacing for the generic MRF structures. It is seen that logarithmic spacing produces a
slight reduction in dispersion compared to arithmetic spacing for most of the structures,
as most of the data points are below the x = y line in Figure 7.20. However, the
difference in dispersion is generally very small, as it is approximately 0.01 lnSaAvg units
or a relative difference of 5%, on average, for the 396 structures.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 260

Figure 7.20: Comparison of the overall minimum dispersion, lnSaAvg(a*,b*), achieved with
logarithmic versus arithmetic period spacing for the 396 generic MRF structures.

While it is observed that the overall minimum dispersion achieved with logarithmic
versus arithmetic spacing is generally comparable, a very important point is that the
values of a* and b* associated with the period range that produces the minimum
dispersion are not the same for both spacing schemes. The average value of a* increased
from 0 for arithmetic spacing to 0.3 for logarithmic spacing. An increase in the b* value
was also observed for logarithmic spacing, with the b* value for logarithmic spacing
being double, on average, that for arithmetic spacing. The average b* value is 3 for
arithmetic spacing and 6 for logarithmic spacing.

It is important to be able to predict a priori a period range that will result in an efficient
estimation (i.e., a small dispersion) of the collapse intensities measured by Saavg for a
given structure because (1) it is generally impractical to compute the dispersion of
collapse intensities for multiple period ranges; and, more importantly, (2) one should
obtain the collapse intensities using ground motions that are appropriate for the site and
the hazard level of interest as identified through a PSHA deaggregation, which is a
function of the period range used to compute Saavg. The overall minimum dispersion
lnSaAvg(a*,b*) obtained via logarithmic period spacing was generally (slightly) smaller than
that obtained via arithmetic spacing (Figure 7.20), but it remains to be seen which period
spacing scheme leads to more efficiency when the period range used to compute Saavg is
predicted as a function of structural properties. Figure 7.21(a) compares the dispersion of
the 396 generic MRF structures computed using the period-independent a and b values
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 261

previously recommended for use with arithmetic and logarithmic spacing. It is seen that
logarithmic spacing generally results in reduced dispersion for most structures, but the
difference is small. The average reduction is 0.02 lnSaAvg units or a relative reduction of
6%. A slight improvement is achieved for arithmetic spacing when the regression seen in
Figure 7.13 is used to determine the b. Figure 7.21(b) compares these dispersions to those
computed using Saavg(0.2, 9) for logarithmic spacing, but it is seen that logarithmic
spacing is still slightly more efficient, on average, than arithmetic spacing. The average
reduction is 0.01 lnSaAvg units or a relative reduction of 5%. It should be noted that the
results presented here are based on sampling spectral values at 100 periods within a given
period range but that when only 10 periods were used, the results did not change
significantly.

Figure 7.21: Comparison of the dispersion, lnSaAvg(a,b), achieved with logarithmic versus
arithmetic period spacing for the 396 generic MRF structures using: (a) Saavg(0.2, 9) for
logarithmic spacing versus Saavg(0, 3) for arithmetic spacing; (b) Saavg(0.2, 9) for
logarithmic spacing versus Saavg(0, b:reg) for arithmetic spacing. Note that b:reg is the b
value obtained from the regression based on T1 shown in Figure 7.13.

7.4.4 Modern reinforced concrete moment-resisting frame structures


Thirty RC MRF structures used by Haselton and Deierlein (2007, Chapter 3) are analyzed
here to develop recommendations for the range of periods (i.e. values of a and b) to use
when computing Saavg and to investigate the effect of period spacing when computing
Saavg. Structures with 1, 2, 4, 8, 12, and 20 stories are included. The fundamental period
of the structures ranges from T1 = 0.42 s to 2.63 s. For a given number of stories the
structures vary in their design decisions (e.g., perimeter frame versus space frame, bay
width, conservatism in column sizes and reinforcement ratios, etc.). All structures were
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 262

analyzed with the Set One ground motion set. Additional details about these structures
are provided in Appendix B.

7.4.4.1 Arithmetic period spacing


Figure 7.22 presents contours of the dispersion of the collapse intensities as a function of
the a and b values used to compute Saavg(a, b) with arithmetic period spacing that are
representative of results for the 1-story structures (Figure 7.22(a)), the 2-story structures
(Figure 7.22(b)), the 4-story structures (Figure 7.22(c)), the 8-story structures (Figure
7.22(d)), the 12-story structures (Figure 7.22(e)), and the 20-story structures (Figure
7.22(f)). As indicated by Figure 7.22 both the a* and b* values generally decrease as the
number of stories increases. The trend of b* decreasing with increasing number of stories
was observed for the generic MRF structures; however, the a* values were generally
close to 0 for structures greater than four stories (Figure 7.11). The a* values of the 4-
story generic MRF structures varied between 0 and 1, but the value of a generally did not
have a large influence on the dispersion of Saavg(a, b), which is also observed for the 1-,
2- and 4-story RC MRFs in Figure 7.22.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 263

Figure 7.22: Contours of the dispersion (ln) of the collapse intensities as a function of the
a and b values used to compute Saavg(a, b) using arithmetic period spacing that are
representative of results for RC MRF structures: (a) 1 story; (b) 2 story; (c) 4 story; (d) 8
story; (e) 12 story; (f) 20 story. Note the different vertical scales between plots (a), (b)
and (c) versus plots (d), (e) and (f).

The trends indicated by Figure 7.22 are better illustrated in Figure 7.23, which shows a
scatter plot of the a* and b* values for each of the 30 RC MRF structures. As previously
indicated, both the a* and b* values generally decrease as the number of stories
increases. Of the four one-story structures considered, only one of them has an a* value
less than 1. This structure has a fundamental period of T = 0.71 s, while the other one-
story structures all have periods of T = 0.42 s. The a* value of the generic SDOF systems
investigated in Section 7.3.2 was less than 1, even for structures with T 0.5 s. Despite
the difference in a* values, the value of a used to compute Saavg(a, b) generally has a
minor influence on the dispersion of the collapse intensities for short period structures.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 264

Figure 7.23: Scatter plot of a* and b* using arithmetic period spacing for all RC MRF
structures.

Figure 7.24 shows the b* values of all 30 RC MRF structures plotted as a function of T1
in the log-log domain and includes a linear regression fit to the data. One-story structures
with T1 = 0.43 s are excluded from the regression as their b* values were approximately
5, and this was not captured by the linear regression. The regression indicates that the b*
value is approximately 6.0 for structures with T1 = 0.5 s and decreases linearly (in the
log-log domain) to approximately 2.7 when T1 = 2.5 s. As seen in Figure 7.24 the
regression fits the data very well, which is reflected by the correlation coefficient of = -
0.97 (or equivalently R2 = 2 = 0.94).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 265

Figure 7.24: b* value associated with the minimum dispersion of Saavg(a, b) as a function
of T1 using arithmetic period spacing for the RC MRF structures.

The regressions of b* versus T1 shown in Figure 7.13 for the generic MRF structures and
in Figure 7.24 for the RC MRF structures are similar, with both regressions predicting a
b* value of approximately 3 for T1 = 2 s. For T1 < 2 the RC MRF regression predicts
slightly higher b* values, although it is based on significantly fewer data points than the
generic MRF regression.

To determine the a and b values to use for a period-independent recommendation (i.e.,


one that is not a function of structural properties such as T1), the difference between the
dispersion on the collapse intensities computed using Saavg(a*, b*) (i.e., the overall
minimum dispersion for a given structure) versus Saavg(a, b) for given values of a and b is
computed for each of the 30 RC MRF structures. The mean value of this difference for all
30 structures is shown in Figure 7.25, which shows a contour plot of the difference in
dispersion as a function of the a and b values used to compute Saavg(a, b). Figure 7.25(a)
presents the results when the difference is computed in absolute terms (e.g., lnSaAvg(a,b) -
lnSaAvg(a*,b*)), and Figure 7.25(b) presents the results when the absolute difference is
normalized by the value of lnSaAvg(a*,b*). The latter case essentially represents a percent
difference and is used because the overall minimum dispersion varies between structures.
Both plots in Figure 7.25 indicate that a = 0 and b = 3.5 are convenient values to use
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 266

when computing Saavg(a, b) as these values are associated with the smallest difference, on
average, from the absolute minimum dispersion that can be calculated for a given
structure using Saavg. The absolute difference is on the order of 0.04 lnSaAvg units when
using a = 0 and b = 3.5 as seen in Figure 7.25(a), and the relative difference is on the
order of 0.24, or 24%, as seen in Figure 7.25 (b). These differences are larger than those
seen for the generic MRF structures and are largely due to the presence of the 1- and 2-
story structures in the RC MRF set, which tend to have higher b* values than structures
with more stories. This also contributes to a slightly higher recommended period-
independent b value for the RC MRF structures compared to the generic MRF structures
(b = 3.5 versus 3, respectively).

Figure 7.25: Contour plot showing the mean value, based on the 30 RC MRF structures,
of the difference between the dispersion on the collapse intensities computed using
Saavg(a*, b*) versus Saavg(a, b) as a function of a and b using arithmetic period spacing:
(a) difference in absolute terms; (b) difference normalized by the dispersion based on
Saavg(a*, b*).

Figure 7.26 is used to investigate how the efficiency (as measured by the dispersion of
the collapse intensities) of Saavg(0, b) is affected when b is calculated using the regression
with T1 shown in Figure 7.24 (denoted as b:reg) versus the period-independent
recommendation of b = 3.5 and how these compare to the efficiency of Saavg(a*, b*).
Figure 7.26(a) is a scatter plot comparing the value of lnSaAvg(a*,b*) to lnSaAvg(0,b:reg) for
each of the 30 RC MRF structures. Note that the b:reg value predicted by the equation
shown in Figure 7.24 is used for structures with T1 = 0.42 s, even though these structures
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 267

were not used in determining the regression coefficients. Many points in Figure 7.26(a)
lie on or near the x = y line, which indicates that the regression is very effective at
predicting a b value that, when used with a = 0, will give a dispersion that is essentially
the same as the overall minimum dispersion for a given structure. This is reinforced by
that fact that the point denoting the mean dispersion for all structures is close to the
x = y line. The close match between lnSaAvg(a*,b*) and lnSaAvg(0,b:reg) is somewhat
expected based on the high correlation coefficient between b* and b:reg shown in Figure
7.24 coupled with the fact that the a value generally does not have a large effect on the
dispersion. The period-independent values of a = 0 and b = 3.5 produce more scatter with
respect to lnSaAvg(a*,b*), as shown in Figure 7.26(b), compared to using a = 0 and b =
b:reg. Figure 7.26(c) compares lnSaAvg(0,3.5) to lnSaAvg(0,b:reg), and it is seen that using a
period-dependent b versus a period-independent value of b = 3.5 generally leads to equal
or smaller dispersion as nearly all points, including the mean point, lie on or above the
x = y line. The absolute difference in the dispersion when b is period-dependent versus
using the period-independent value of b = 3.5 is generally small; however, for nearly a
third of the RC MRF structures using the period-independent value of b = 3.5 increases
the dispersion by at least 25% compared to the value computed using b:reg.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 268

Figure 7.26: Scatter plot of dispersion in collapse intensities computed using Saavg for the
30 RC MRF structures for different computations of Saavg(a, b) using arithmetic period
spacing: (a) Saavg(a*, b*) vs. Saavg(0, b:reg); (b) Saavg(a*, b*) vs. Saavg(0, 3.5); (c)
Saavg(0, 3.5) vs. Saavg(0, b:reg). Note that b:reg is the b value obtained from the
regression based on T1 shown in Figure 7.24.

7.4.4.2 Logarithmic period spacing


Figure 7.27 presents contours of the dispersion of the collapse intensities as a function of
the a and b values used to compute Saavg(a, b) with logarithmic period spacing that are
representative of results for the 1-story structures (Figure 7.27(a)), the 2-story structures
(Figure 7.27(b)), the 4-story structures (Figure 7.27(c)), the 8-story structures (Figure
7.27(d)), the 12-story structures (Figure 7.27(e)), and the 20-story structures (Figure 7.27
(f)). As indicated by Figure 7.27 both the a* and b* values generally decrease as the
number of stories increases. The trend of b* decreasing with increasing number of stories
was also observed for the generic MRF structures.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 269

Figure 7.27: Contours of the dispersion (ln) of the collapse intensities as a function of the
a and b values used to compute Saavg(a, b) using logarithmic period spacing that are
representative of results for RC MRF structures: (a) 1 story; (b) 2 story; (c) 4 story; (d) 8
story; (e) 12 story; (f) 20 story. Note the different vertical scales between plots (a), (b)
and (c) versus plots (d), (e) and (f).

The relationship of the a* and b* values with T1 are shown in Figure 7.28(a) and Figure
7.28(b), respectively. Both the a* and b* show a strong linear trend with T1 in the log-log
domain, as the correlation coefficients are = -0.7 and -0.8, respectively. The strong
relationship between a* and T1 was not observed for the generic MRF structures when
using logarithmic spacing.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 270

Figure 7.28: Relationship between the a* and b* values associated with the minimum
dispersion of Saavg(a, b) obtained using logarithmic period spacing and T1 for the 30 RC
MRF structures: (a) a* and T1; (b) b* and T1.

Figure 7.29 is used to investigate how the efficiency (as measured by the dispersion of
the collapse intensities) of Saavg(a, b) is affected when a and b are calculated using the
regression with T1 shown in Figure 7.28 (denoted as a:reg and b:reg, respectively) versus
the period-independent recommendation of a = 0.4 and b = 9 and how these compare to
the efficiency of Saavg(a*, b*). The values of a and b used for the period-independent
recommendation are the values that, on average based on the 30 RC MRF structures,
result in the smallest difference between the overall minimum dispersion of a given
structure and the dispersion calculated using period-independent values of a and b. As
before, a maximum period of T = 10 s was used to compute Saavg even if bT1 exceeded
10 s (i.e., b for the period-independent recommendation was taken as the minimum of 9
and 10/T1). Figure 7.29(a) is a scatter plot comparing the value of lnSaAvg(a*,b*) to
lnSaAvg(a:reg,b:reg) while Figure 7.29(b) compares lnSaAvg(a*,b*) to lnSaAvg(0.4,9) and Figure
7.29(c) lnSaAvg(0.4,9) to lnSaAvg(a:reg,b:reg). It can be seen in Figure 7.29(a) that using the
regression-based values of a and b produces dispersions that are a very close match to the
overall minimum dispersion. This is not unexpected as the regressions predicted a and b
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 271

values that were generally close to the a* and b* values, respectively as seen in Figure
7.28. The period-independent values of a = 0.2 and b = 9 generally produce a dispersion
close to the overall minimum dispersion of a given structure as seen in Figure 7.29(b);
however, for some structures using the difference can be large. Figure 7.19(c) shows that
the regression-based (i.e., period-dependent) values of a and b will generally lead to
smaller dispersions than the period-independent values of a = 0.2 and b = 9. On average,
there is a 9% reduction in dispersion using the regression-based values compared to the
period-independent values, which is equivalent to an absolute reduction of 0.02 lnSaAvg
units.

Figure 7.29: Comparison of the dispersion, lnSaAvg(a,b), achieved with logarithmic period
spacing for the 30 RC MRF structures for different computations of Saavg(a, b) using
arithmetic period spacing: (a) Saavg(a*, b*) vs. Saavg(a:reg, b:reg); (b) Saavg(a*, b*) vs.
Saavg(0.4, 9); (c) Saavg(0.4, 9) vs. Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a
and b values obtained from the regressions based on T1 shown in Figure 7.28.

7.4.4.3 Comparison of period spacing schemes


Figure 7.30 compares the dispersion of collapse intensities computed using Saavg with
arithmetic versus logarithmic period spacing schemes for the 30 RC MRF structures.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 272

Figure 7.30(a) compares the overall minimum dispersion lnSaAvg(a*,b*), and it is observed
that there is not a significant difference in the value computed using arithmetic versus
logarithmic period spacing. Figure 7.30(b) compares the dispersion computed using the
period-independent a and b values previously recommended for use with arithmetic and
logarithmic spacing (i.e., Saavg(0.4, 9) for logarithmic spacing versus Saavg(0, 3.5) for
arithmetic spacing). It is observed that logarithmic spacing generally results in a slight
reduction in dispersion, but the difference is relatively small (an average reduction of 3%,
or less than 0.01 lnSaAvg units). Figure 7.30(c) compares the dispersion computed using
Saavg(0, b:reg) for arithmetic spacing (where b:reg is determined from Figure 7.24) and
Saavg(a:reg, b:reg) for logarithmic spacing (where a:reg and b:reg are determined from
Figure 7.28), and it is observed that there is generally not a significant difference in the
dispersions computed using arithmetic versus logarithmic period spacing when using
period-dependent values of a and b. It should be noted that the results presented here are
based on sampling spectral values at 100 periods within a given period range but that
when only 10 periods were used, the results did not change significantly.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 273

Figure 7.30: Comparison of the dispersion, lnSaAvg(a,b), achieved with logarithmic versus
arithmetic period spacing for the 30 RC MRF structures using: (a) the overall minimum
dispersion; (b) period-independent (rigid) recommendations of Saavg(0.4, 9) for
logarithmic spacing versus Saavg(0, 3.5) for arithmetic spacing; (c) regression-based
values of Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a and b values,
respectively, obtained from the regressions based on T1 shown in Figure 7.24 for
arithmetic spacing and in Figure 7.28 for logarithmic spacing. Also note that for
arithmetic spacing a:reg = 0.

7.4.5 Generic shear wall structures


Approximately 250 generic shear wall structures used by Zareian and Krawinkler (2009)
are analyzed here to again develop recommendations for the range of periods (i.e. values
of a and b) to use when computing Saavg in these types of structures and to investigate the
effect of period spacing when computing Saavg. The behavior of these structures is
governed by flexure, and shear deformations and shear failure are not considered.
Structures of 4, 8, 12 and 16 stories are included, and the first mode period of the
structure for a given number of stories (N) is either T1 = 0.05N, 0.075N or 0.1N
seconds. Structural parameters including the ductility-dependent strength reduction factor
(R) and the post-elastic behavior of components are also varied. All structures were
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 274

analyzed with the LMSR-N ground motion set. Additional details about these structures
are provided in Appendix B.

7.4.5.1 Arithmetic period spacing


Figure 7.31 presents contours of the dispersion of the collapse intensities as a function of
the a and b values used to compute Saavg(a, b) with arithmetic period spacing that are
representative of results for the 4-story structures (Figure 7.31(a)), the 8-story structures
(Figure 7.31(b)), the 12-story structures (Figure 7.31(c)), and the 16-story structures
(Figure 7.31(d)). As indicated by Figure 7.31 a* is generally 0, which indicates the
minimum dispersion is achieved when the period range starts from T = 0. The 4-story
structures are generally an exception to this statement; however, the a value typically has
minimal effect on the dispersion of Saavg(a, b) as indicated by Figure 7.31(a). Figure 7.31
also indicates that the value of b* generally decreases as the number of stories increases,
which was also observed for the MRF structures.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 275

Figure 7.31: Contours of the dispersion (ln) of the collapse intensities as a function of the
a and b values used to compute Saavg(a, b) using arithmetic period spacing that are
representative of results for generic shear wall structures: (a) 4 story; (b) 8 story; (c) 12
story; (d) 16 story. Note the different vertical scales between plots (a) and (b) versus plots
(c) and (d).

The trends implied by Figure 7.31 are better illustrated in Figure 7.32, which shows
histograms representing the joint distribution of the a* and b* values for the 252
structures. The histograms are separated based on number of stories, and for a given
number of stories there are 63 structures. As previously noted, the 4-story structures
shown in Figure 7.32(a) tend to have more scatter in the a* values than the other
structures and significantly more scatter in the a* values than the 4-story generic MRF
structures presented in Figure 7.11(a). With the exception of the 4-story structures, the
vast majority a* values for the generic shear walls structures in Figure 7.32 tend to be
less than 0.10. The histograms in Figure 7.32 also show that the b* values tend to be
concentrated between 4 and 10, and the overall scatter in a* and b* values tends to
decrease as the number of stories increases.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 276

Figure 7.32: Histogram of a* and b* values associated with the minimum dispersion of
Saavg(a, b) for generic shear wall structures using arithmetic period spacing: (a) 4 story;
(b) 8 story; (c) 12 story; (d) 16 story. Note the different axes used for (a).

Figure 7.33 shows the b* values for each of the 252 structures. The horizontal axis in
Figure 7.33 that displays the structure number (a value between 1 and 252) was added to
make the data easier to see because the points are tightly clustered in some regions. The
vertical lines divide the structures into bins based on fundamental period. For a given
number of stories, the left, center, and right bins indicate structures with fundamental
periods of T1 = 0.05N, 0.075N or 0.1N seconds, respectively. Figure 7.33 shows that
the value of b* for a given number of stories tends to decrease as the fundamental period
of the structure increases, and this trend generally becomes more pronounced as the
number of stories increases. A similar trend was observed for the generic MRF structures.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 277

Figure 7.33: Scatter plot of b* for all generic shear wall structures using arithmetic period
spacing.

Figure 7.34 shows the b* values of all 252 generic shear wall structures plotted as a
function of T1 in the log-log domain. A linear regression is also fit to the data and
indicates that the b* value is approximately 11 for structures with T1 = 0.2 s and
decreases to approximately 4 when T1 = 1.6 s. As seen in Figure 7.34 there is a
considerable amount of scatter about the regression, which is reflected by the correlation
coefficient of = -0.65 (or, equivalently, R2 = 2 = 0.42). Compared to the generic MRF
structures, the generic shear wall structures generally have a higher b* value for a given
value of T1, although the shear walls structures typically have a greater number of stories
than the MRF structures for a given T1.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 278

Figure 7.34: b* value associated with the minimum dispersion of Saavg(a, b) using
arithmetic period spacing as a function of T1 for the generic shear wall structures.

Other trends between the value of b* and structural properties (e.g., strength and post-
yield structural component parameters) are examined. Though the relationship between
b* and structural properties appears to be most significantly influenced by the
fundamental period of the structure, a trend between R and b* is also observed. For a
given number of stories and a given fundamental period, the b* value generally increases
with increasing values of R. In particular, it was observed that systems with R = 1.5
could have b* values up to 50% less, on average, than systems with R = 3.0 or 6.0 (the
other two values of R that were used for the generic shear wall structures). The disparity
in b* values is most notable for the 8-story structures, which can be observed in Figure
7.33 as there are generally two distinct clusters of b* values within a period bin for the 8-
story structures. The disparity in b* values decreases as number of stories increases,
which is also reflected in Figure 7.33 as the variability of b* values within a period bin
decreases with increasing number of stories.

To develop a general recommendation of period-independent a and b values to use for


computing Saavg, the difference between the dispersion on the collapse intensities
computed using Saavg(a*, b*) (i.e., the overall minimum dispersion for a given structure)
versus Saavg(a, b) for given values of a and b is computed for each of the 252 generic
shear wall structures. The mean value of this difference for all 252 structures is shown in
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 279

Figure 7.35, which shows a contour plot of the difference in dispersion as a function of
the a and b values used to compute Saavg(a, b). Figure 7.35(a) presents the results when
the difference is computed in absolute terms (e.g., lnSaAvg(a,b) - lnSaAvg(a*,b*)), and Figure
7.35(b) presents the results when the absolute difference is normalized by the value of
lnSaAvg(a*,b*). The latter case essentially represents a percent difference and is used
because the overall minimum dispersion varies between structures from lnSaAvg(a*,b*) =
0.09 for one of the 4-story structures to lnSaAvg(a*,b*) = 0.25 for one of the 16-story
structures. Both plots in Figure 7.35 indicate that a = 0 and b = 6 are convenient values to
use when computing Saavg(a, b) as these values show the smallest difference, on average,
from the absolute minimum dispersion that can be calculated for a given structure using
Saavg. The absolute difference is on the order of 0.05 lnSaAvg units when using a = 0 and b
= 6 as seen in Figure 7.35(a), and the relative difference is on the order of 0.3, or 30%, as
seen in Figure 7.35(b).

Figure 7.35: Contour plot showing the mean value, based on the 252 generic shear wall
structures, of the difference between the dispersion on the collapse intensities computed
using Saavg(a*, b*) versus Saavg(a, b) as a function of a and b for arithmetic period
spacing: (a) difference in absolute terms; (b) difference normalized by the dispersion
based on Saavg(a*, b*).

Figure 7.36 is used to investigate how the efficiency (as measured by the dispersion of
the collapse intensities) of Saavg(0, b) is affected when b is calculated using the regression
with T1 shown in Figure 7.34 (denoted as b:reg) versus the period-independent
recommendation of b = 6 and how these compare to the efficiency of Saavg(a*, b*).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 280

Figure 7.36(a) is a scatter plot comparing the value of lnSaAvg(a*,b*) to lnSaAvg(0,b:reg) for
each of the 252 generic shear wall structures. Many points in Figure 7.36(a) lie on or near
the x = y line, which indicates that the regression is generally effective at predicting a b
value that, when used with a = 0, will give a dispersion that is essentially the same to the
overall minimum dispersion for a given structure. This is reinforced by that fact that the
point denoting the mean dispersion for all structures is close to the x = y line. Some
points, particularly those where lnSaAvg(a*,b*) 0.1, lie far from the x = y line, which
may be expected due to the scatter between b* and b:reg shown in Figure 7.34. These
structures tend to be 4-story structures with the least ductile structural components and
the smallest R factors. The period-independent values of a = 0 and b = 6 produce more
scatter with respect to lnSaAvg(a*,b*), as shown in Figure 7.36(b), compared to using a = 0
and b = b:reg. Figure 7.36(c) compares lnSaAvg(0,3.5) to lnSaAvg(0,b:reg), and it is seen that
computing b using the regression versus a period-independent value of b = 6 generally
leads to equal or smaller dispersion as most points, including the mean point, lie on or
above the x = y line. The absolute difference in the dispersion when b is computed
from the regression versus using the period-independent value of b = 6 is generally small;
however, for nearly a fifth of the generic shear wall structures using the period-
independent value of b = 6 increases the dispersion by at least 25% compared to the value
computed using b:reg. These structures are generally the same 4-story structures with the
least ductile structural components and the smallest R factors for which the value of
lnSaAvg(0,b:reg) is significantly larger than lnSaAvg(a*,b*).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 281

Figure 7.36: Scatter plot of dispersion in collapse intensities computed using Saavg with
arithmetic period spacing for the 252 generic shear wall structures for different
computations of Saavg(a, b): (a) Saavg(a*, b*) vs. Saavg(0, b:reg); (b) Saavg(a*, b*) vs.
Saavg(0, 6); (c) Saavg(0, 6) vs. Saavg(0, b:reg). Note that b:reg is the b value obtained from
the regression based on T1 shown in Figure 7.34.

7.4.5.2 Logarithmic period spacing


Figure 7.37 presents contours of the dispersion of the collapse intensities as a function of
the a and b values used to compute Saavg(a, b) with logarithmic period spacing that are
representative of results for the 4-story structures (Figure 7.37(a)), the 8-story structures
(Figure 7.37(b)), the 12-story structures (Figure 7.37(c)), and the 16-story structures
(Figure 7.37(d)). As indicated by Figure 7.37(a) a* is generally 1 for the 4-story
structures and also for some of the 8-story structures. It was investigated whether
considering a values greater than 1 (i.e., only using periods greater than the fundamental
period of the system) would significantly reduce the dispersion relative to what was
found using a 1. Based on a sample of structures with a* = 1 (when a was constrained
to be 1), it was found that allowing a* to be greater than one did not generally result in
a significant reduction in dispersion. Figure 7.37 also indicates that the value of b*
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 282

generally decreases as the number of stories increases, which was also observed for the
MRF structures.

Figure 7.37: Contours of the dispersion (ln) of the collapse intensities as a function of the
a and b values used to compute Saavg(a, b) with logarithmic period spacing that are
representative of results for generic shear wall structures: (a) 4 story; (b) 8 story;
(c) 12 story; (d) 16 story.

The relationship of the a* and b* values with T1 are shown in Figure 7.38(a) and Figure
7.38(b), respectively. Both the a* and b* show a significant linear trend with T1 in the
log-log domain, as the correlation coefficients are approximately = -0.7 and -0.6,
respectively. Note that structures with T1 0.4 s are excluded from the regression
between a* and T1 as nearly all of these structures have a* = 1.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 283

Figure 7.38: Relationship between the a* and b* values associated with the minimum
dispersion of Saavg(a, b) obtained using logarithmic period spacing and T1 for the 252
generic shear wall structures: (a) a* and T1; (b) b* and T1.

Figure 7.39 is used to investigate how the efficiency (as measured by the dispersion of
the collapse intensities) of Saavg(a, b) is affected when a and b are calculated using the
regression with T1 shown in Figure 7.38 (denoted as a:reg and b:reg, respectively) versus
the period-independent recommendation of a = 0.7 and b = 15 and how these compare to
the efficiency of Saavg(a*, b*). The values of a and b used for the period-independent
recommendation are the values that, on average based on the 252 generic shear wall
structures, result in the smallest difference between the overall minimum dispersion of a
given structure and the dispersion calculated using period-independent values of a and b.
Note that a maximum period of T = 10 s was used to compute Saavg even if bT1 exceeded
10 s (i.e., b for the period-independent recommendation was taken as the minimum of 15
and 10/T1), which means that for structures with T1 0.67 s Saavg is computed using
periods between 0.7T1 and 10 s. Figure 7.39(a) is a scatter plot comparing the value of
lnSaAvg(a*,b*) to lnSaAvg(a:reg,b:reg) while Figure 7.39(b) compares lnSaAvg(a*,b*) to lnSaAvg(0.7,15)
and Figure 7.39(c) compares lnSaAvg(0.7,15) to lnSaAvg(a:reg,b:reg). The dispersions computed
using the regression-based values of a and b are generally close to the overall minimum
dispersions; however, for some structures where lnSaAvg(a*,b*) 0.1, the regression-based
dispersions are significantly higher than the overall minimum as seen in Figure 7.39(a).
These structures tend to be 4-story structures with the least ductile structural components
and the smallest R factors, and observation was made regarding these structures when
considering arithmetic period spacing. The period-independent values of a = 0.7 and
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 284

b = 15 produce more scatter with respect to lnSaAvg(a*,b*), as shown in Figure 7.39(b),


compared to regression based-values. Figure 7.39(c) directly compares the dispersion
computed using the regression-based and period-independent a and b values, and it is
observed that the regression-based values generally lead to smaller dispersions, with an
average reduction of 6%.

Figure 7.39: Scatter plot of the dispersion in collapse intensities computed using Saavg
with logarithmic period spacing for the 252 generic shear wall structures for different
computations of Saavg(a, b): (a) Saavg(a*, b*) vs. Saavg(a:reg, b:reg); (b) Saavg(a*, b*) vs.
Saavg(0.7,15); (c) Saavg(0.7,15) vs. Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a
and b values, respectively, obtained from the regression based on T1 shown in Figure
7.38.

7.4.5.3 Comparison of period spacing schemes


Figure 7.40 compares the dispersion of collapse intensities computed using Saavg with
arithmetic versus logarithmic period spacing schemes for the 252 generic shear wall
structures. Figure 7.40(a) compares the overall minimum dispersion lnSaAvg(a*,b*), and it is
observed that there is generally not a significant difference in the value computed using
arithmetic versus logarithmic period spacing. Figure 7.40(b) compares the dispersion
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 285

computed using the period-independent a and b values previously recommended for use
with arithmetic and logarithmic spacing (i.e., Saavg(0.7, 15) for logarithmic spacing
versus Saavg(0, 6) for arithmetic spacing). It is observed that difference in the dispersions
computed using arithmetic versus logarithmic period spacing is not significant, on
average. The dispersions computed using the period-independent a and b values (for
either period spacing scheme) are generally higher than those computed using the
regression-based values, which are shown in Figure 7.40(c). The dispersions computed
using the regression-based values are generally not significantly different for logarithmic
versus arithmetic period spacing. It should be noted that the results presented here are
based on sampling spectral values at 100 periods within a given period range but that
when only 10 periods were used, the results did not change significantly.

Figure 7.40: Comparison of the dispersion, lnSaAvg(a,b), achieved with logarithmic versus
arithmetic period spacing for the 252 generic shear wall structures using: (a) the overall
minimum dispersion; (b) period-independent (rigid) recommendations of Saavg(0.7, 15)
for logarithmic spacing versus Saavg(0, 6) for arithmetic spacing; (c) regression-based
values of Saavg(a:reg, b:reg). Note that a:reg and b:reg are the a and b values,
respectively, obtained from the regressions based on T1 shown in Figure 7.34 for
arithmetic spacing and in Figure 7.38 for logarithmic spacing. Also note that for
arithmetic spacing a:reg = 0.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 286

7.5 Summary and discussion


This chapter explored how the range, number, and spacing (i.e., arithmetic versus
logarithmic) of periods used to compute Saavg influences the effectiveness of this
intensity measure (IM), with a primary focus on the impact these variables have on the
efficiency of the IM, which is measured by the lognormal standard deviation, or
dispersion, of the collapse intensities.

The effect of period range on the efficiency of Saavg was investigated using collapse data
from a large number of structures including generic SDOF systems, generic moment-
resisting frame (MRF) systems, reinforced concrete (RC) MRF systems, and generic
shear wall systems. Let Saavg(a, b) denote the value of Saavg computed using a period
range between aT1 and bT1, where a and b are non-negative constants such that a b
and T1 is the fundamental period of the structure. In these studies, values of a between 0
and 1 and values of b between 1 and 10/T1 were considered (i.e., the lower limit of the
period range was between 0 and T1, and the upper limit was between T1 and 10 s). A limit
of 10/T1 was placed on the b value for practical reasons, primarily due to a general lack of
ground motion prediction equations that predict spectral values at periods longer than
10 s that permit the computation of seismic hazard data needed for collapse (and other
structural response-related) risk assessments and because spectral ordinates at periods
longer than 10 s are typically strongly affected by baseline correction, high-pass filtering
and other record processing modifications applied to ground motion recordings.

For the vast majority of structures examined, the efficiency of Saavg was much more
sensitive to the b value (i.e., the upper limit of the period range) than to the a value (i.e.,
the lower limit of the period range). Let a* and b* be the optimum values of a and b,
respectively, such that Saavg(a*, b*) produces the minimum dispersion in the collapse
intensities in a given structure under a given set of ground motions. It was observed that
the values of a* and b* or, more generally, the period ranges that tended to produce the
smallest dispersions in collapse intensities were primarily a function of the structural
system (i.e., MRF versus shear wall) and of the fundamental period of the system. Other
structural properties such as strength and ductility were also observed to affect the period
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 287

range in some cases; however, these properties had a much smaller effect on the values of
a* and b* than changes in the structural system or in the fundamental period.

The dispersion of collapse intensities measured by Saavg using arithmetically spaced


periods was investigated for a set of generic, relatively ductile SDOF systems with
periods between T = 0.1 s and 4.0 s. It was found that there was a benefit to including
periods less than the fundamental period in the period range used to compute Saavg as the
minimum dispersion is achieved at a values less than 1 for all the systems examined. This
may be a counterintuitive result for those who subscribe to the hypothesis that accounting
for spectral ordinates at periods shorter than the fundamental period is necessary to
account for higher mode effects, and therefore for an SDOF system there would be no
need or benefit to including periods to the left of T1 when computing Saavg; however,
results clearly indicate that including information about spectral ordinates at periods
smaller than the fundamental period leads to reductions in dispersions even for SDOF
systems (which do not have higher modes of vibration). As discussed in Section 5.6, it is
believed that the reduction in dispersion seen when including periods less the
fundamental period in the spectral averaging range results from the fact that the response
spectrum is a reflection of acceleration pulses in the ground motion, and even pulses that
manifest themselves in the spectrum at periods less than the fundamental period can
affect the nonlinear response of the system. As discussed in Chapter 5, using an average
spectral value also minimizes the role that the initial conditions play in the response of
the elastic system, which is important as a nonlinear system will, in general, not have the
same initial conditions as the elastic system and can therefore respond very differently
than the elastic system to the same ground motion segment. For this reason, it makes
sense to take the average spectral value using periods smaller and larger than the
fundamental period as neighboring systems (i.e., systems with periods similar to the
fundamental period) can represent how the elastic system may have responded given
different initial conditions. Additional details regarding this reasoning are provided in
Sections 5.4 and 5.5.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 288

A large number of structures used by previous investigators were analyzed to develop


recommendations for the range of periods (i.e. values of a and b) to use when computing
Saavg in these types of structures and to investigate the effect of period spacing when
computing Saavg. Structures used in this section include 396 generic MRF structures and
252 generic shear wall structures analyzed by Zareian and Krawinkler (2009, Chapter 4)
and 30 modern RC MRF structures analyzed by Haselton and Deierlein (2007, Table 3-
3). The generic MRF and generic shear wall structures are either 4, 8, 12, or 16 stories,
and the fundamental periods of these structures are a function of the number of stories
(0.10N, 0.15N, or 0.20N for the MRF structures and 0.005N, 0.075N, or 0.100N for
the shear wall structures) and vary from T1 = 0.4 s to 3.2 s for the MRF structures and
from T1 = 0.2 s to 1.6 s for the shear wall structures. The shear walls are governed by
flexural behavior and assumed to fail in flexure, not shear. The 30 modern RC MRF
structures vary from 1 to 20 stories with fundamental periods ranging from T1 = 0.42 s to
2.63 s. Additional information about these structures is provided in Appendix B and in
the sources previously cited.

The average value of the minimum dispersion in collapse intensities achieved for a given
structure was approximately lnSaAvg(a*,b*) = 0.20 for the generic MRF structures and
lnSaAvg(a*,b*) = 0.16 for both the RC MRF and the generic shear wall structures. The use of
Saavg when computed using an optimum range of periods Saavg(a*, b*) led to significant
reductions in the dispersion of collapse intensities compared to those computed using
Sa(T1): the average reduction was nearly 50% for the generic MRF structures and over
60% for the RC MRF and generic shear wall structures. For reference, the average
dispersion of the collapse intensities computed using Sa(T1) was lnSa(T1) = 0.38 for the
generic MRF structures, lnSa(T1) = 0.44 for the RC MRF structures, and lnSa(T1) = 0.48 for
the generic shear wall structures.

Overall, it was found that the value of lnSaAvg(a*,b*) for a given structure was generally the
same whether arithmetically or logarithmically spaced periods were used to compute
Saavg; however, the values of a* and b* (i.e., the period ranges) that produced the
minimum dispersion, or more generally, the period ranges that resulted in dispersions
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 289

relatively close to the minimum dispersion, were significantly different for arithmetic
versus logarithmic spacing. The following paragraphs discuss trends in the dispersion as
a function of period range that were observed for different structures and for different
period spacing schemes.

When using arithmetic period spacing, it was observed that the dispersion of collapse
intensities was more affected by changes in the b value (i.e., the upper limit of the period
range) than the a value used to compute Saavg. It was also observed that the optimum
value of b* associated with the minimum dispersion for a given structure was negatively
correlated with the fundamental period. In the log-log domain the correlation coefficient
was on the order of = -0.6 and nearly -1.0 for the generic MRF and RC MRF structures,
respectively, for T1 > 0.42 s and on the order of = -0.7 for the generic shear wall
structures. The increased correlation for the RC MRF structures is thought to be the result
of the similarity in design procedures used for these buildings (all were designed using
the same building codes and standards) versus the generic MRF and shear wall structures,
which were developed to investigate the sensitivity of building response to structural
parameters (e.g., strength and structural component properties) and encompass a wide
range of structural parameter values. Linear regressions were used to develop period-
dependent b values by fitting a straight line to the relationship between b* and T1 in the
log-log domain. For the generic MRF structures, the period-dependent value of b
predicted by the regression decreases from approximately 4.1 to 2.5 as the period
increases from T1 = 0.6 s to 3.2 s. For the RC MRF structures, the period-dependent value
of b predicted by the regression decreases from approximately 5.7 to 2.6 as the period
increases from T1 = 0.56 s to 2.63 s. The regressions for the generic and RC MRF
structures intersect near T1 = 2 s, where b is approximately 3. For the generic shear wall
structures, the period-dependent value of b predicted by the regression decreases from
approximately 10.9 to 4.0 as the period increases from T1 = 0.2 s to 1.6 s. For comparison
purposes, the values of b predicted by the regressions at T1 = 1.0 s are approximately 3.5,
4.3, and 5.1 for the generic MRF, RC MRF, and generic shear wall structures,
respectively.
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 290

When Saavg was computed using arithmetic period spacing with a = 0 and period-
dependent b values determined by a linear regression of b* and T1 in the log-log domain,
it was found that the resulting dispersion in collapse intensities was generally very close
to lnSaAvg(a*,b*), the minimum dispersion computed for a given structure using the
optimum period range, with an average difference between 0.01 and 0.03 lnSaAvg units for
the different structure sets. The a value of 0 was chosen as it minimized the average
difference between the overall minimum dispersion, lnSaAvg(a*,b*), and the computed
dispersion, lnSaAvg(0,b:breg); however, the dispersion was generally not very sensitive to the
a value and similar results were achieved using a = 0.2. As an alternative to computing
period-dependent b values based on a regression with T1, a period-independent value of b
was considered (i.e., one that was constant for a given set of structures). It was found that
period-independent values of b = 3 for the generic MRFs, b = 3.5 for the RC MRFs, and
b = 6 for the generic shear walls generally minimized the average difference between the
computed dispersions and the overall minimum dispersion. It was seen that the average
difference between the two dispersions was generally very small, as the average
difference was between 0.02 and 0.04 lnSaAvg units for the different structure sets. In
terms of the difference between the dispersions computed using the period-dependent
versus the period-independent b value, it was found that the period-dependent b value
generally leads to slightly smaller dispersion values, with the average difference less than
0.01 lnSaAvg units for the generic MRF structures, 0.02 lnSaAvg units for the generic shear
wall structures, and 0.03 lnSaAvg units for the RC MRF structures.

When using logarithmic period spacing, it was observed that the dispersion of collapse
intensities was sensitive to both the a and the b values (i.e., both the lower and upper
limits of the period range) used to compute Saavg. Both the optimum a* and b* values
associated with the minimum dispersion for a given structure were negatively correlated
with the fundamental period. In the log-log domain the correlation coefficient between b*
and T1 ranged from approximately = -0.6 for the generic shear wall structures to
approximately = -0.8 for the RC MRF structures, with the generic MRFs falling within
this range. Linear regressions were also used to develop period-dependent b values by
fitting a straight line to the relationship between b* and T1 in the log-log domain. For the
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 291

generic MRF structures, the period-dependent value of b predicted by the regression


decreases from approximately 8.9 to 3.7 as the period increases from T1 = 0.6 s to 3.2 s.
For the RC MRF structures, the period-dependent value of b predicted by the regression
decreases from approximately 8.3 to 3.8 as the period increases from T1 = 0.42 s to
2.63 s. The regressions for the generic and RC MRF structures do not intersect within the
period ranges used for regression; however, the difference in values decreases as T1
increases, with b for the generic MRF structures being approximately 25% larger at T1 =
0.6 s and approximately 10% larger at T1 = 2.63 s. For the generic shear wall structures,
the period-dependent value of b predicted by the regression decreases from
approximately 19.2 to 5.9 as the period increases from T1 = 0.2 s to 1.6 s. For comparison
purposes, the period-dependent values of b predicted by the regressions at T1 = 1.0 s are
approximately 6.8, 5.7, and 7.7 for the generic MRF, RC MRF, and generic shear wall
structures, respectively.

As previously mentioned, the values of a* are also negatively correlated with T1 when
using logarithmic period spacing. The correlation coefficient between a* and T1 in the
log-log domain is approximately = -0.7 for the RC MRF structures and for the generic
shear wall structures with T1 0.6 s but was only on the order of = -0.2 for the generic
MRF structures with T1 0.6 s. For the generic MRF structures, the period-dependent
value of a predicted by the regression decreases from approximately 0.3 to 0.2 as the
period increases from T1 = 0.4 s to 3.2 s. For the RC MRF structures, the period-
dependent value of a predicted by the regression decreases from approximately 0.9 to 0.3
as the period increases from T1 = 0.56 s to 2.63 s. For the generic shear wall structures,
the period-dependent value of a predicted by the regression decreases from
approximately 0.9 to 0.3 as the period increases from T1 = 0.6 s to 1.6 s. For comparison
purposes, the period-dependent values of a predicted by the regressions at T1 = 1.0 s are
approximately 0.2, 0.6, and 0.5 for the generic MRF, RC MRF, and generic shear wall
structures, respectively.

When Saavg was computed using logarithmic period spacing with the period range
defined by period-dependent a and b values, it was found that the resulting dispersions in
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 292

collapse intensities were generally very close to lnSaAvg(a*,b*), the minimum dispersion
computed for a given structure using the optimum period range, with an average
difference between 0.01 and 0.02 lnSaAvg units for the different structure sets. As an
alternative to computing period-dependent a and b values, period-independent values of a
and b were considered (i.e., ones that were constant for a given set of structures). It was
found that the period-independent (a, b) pairs that generally minimized the average
difference between the computed dispersions and the overall minimum dispersion when
using logarithmic period spacing were (0.2, 9) for the generic MRF structures, (0.4, 9) for
the RC MRF structures, and (0.7, 15) for the generic shear wall structures. As previously
noted, periods longer than 10 s were not used to compute Saavg even if bT1 exceeded
10 s. This means that the b value determined by the regression was only used for generic
MRF and RC MRF structures with T1 1.1 s (approximately one-third of the structures in
each set) and for generic shear wall structures with T1 1.5 s (all structures except the 16-
story structures with T1 = 1.6 s); otherwise, the upper limit of the period range used to
compute Saavg was 10 s. The dispersion computed using period-independent values of a
and b was generally close to overall minimum dispersion of a given structure, with an
average difference between 0.01 and 0.04 lnSaAvg units for the different structure sets. In
terms of the difference between the dispersions computed using period-dependent versus
period-independent values of a and b, it was found that using period-dependent values
generally leads to slightly smaller dispersion values, with the average difference being
negligible for the generic MRF structures and equal to approximately 0.01 lnSaAvg units
for the generic shear wall structures and 0.02 lnSaAvg units for the RC MRF structures.

The optimum values of a and b that led to the minimum dispersion in the collapse
intensities were significantly higher for logarithmic spacing than for arithmetic spacing.
This results from the fact that logarithmic spacing samples more heavily from the lower
half of the period range. The increase in the a value for logarithmic spacing is thought to
be the result of the fact that for medium to long period structures, the parts of the ground
motion that control the response of short-period elastic SDOF systems (and produce the
short-period spectral ordinates) are generally not the same as those controlling the
response of longer period systems. Thus, there is a need to avoid over sampling in this
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 293

period region, which is why the value of a increases from the value of 0 that produced
small dispersions when using arithmetic period spacing to values of at least 0.2 for
logarithmic period spacing. In fact, it was often observed that including the very short
period ordinates (T < 0.1T1) could have a detrimental effect on the dispersion when
using logarithmic period spacing. The increase in the b value may be the related to the
increase in a, as increasing the b value for a given a value will reduce the concentration
of samples at shorter periods.

It is observed that when using the previously recommended period-independent values of


a and b (which depend on the period spacing scheme and structure class) to compute
Saavg, logarithmic spacing generally results in a decreased dispersion in the collapse
intensities compared to arithmetic period spacing; however, the difference is, on average,
very small. For the generic MRF structures, the recommended period-independent values
of a and b led to dispersions that were 0.02 lnSaAvg units greater, on average, for
arithmetic spacing compared to logarithmic spacing and 0.01 lnSaAvg units greater, on
average, for the RC MRF and generic shear walls structures.

Given that the differences in the dispersion due to period spacing scheme are, on average,
very small when using either the period-independent or period-independent
recommended values of a and b (which depend on the period spacing scheme and
structure class), one might prefer to use arithmetic period spacing because the
recommended b value for a given structure class is generally much lower for arithmetic
spacing versus logarithmic spacing. Using a lower b value reduces potential issues
associated with a lack of seismic hazard data at long periods, the maximum usable period
of a record, and record processing effects that may significantly affect the values of
spectral ordinates at long periods. The dispersion computed with arithmetic period
spacing is also primarily a function of only the b value whereas the dispersion computed
with logarithmic period spacing can be sensitive to both the a and the b value,
particularly when either value is less than a* or b*, respectively (the a or b value,
respectively, associated with the minimum dispersion of a given structure).
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 294

All results presented here are based on computing Saavg using 100 points/periods within a
given period range. The sensitivity of the results to the number of periods was studied,
and it was found that the results were not significantly affected when only 10
points/periods were used to compute Saavg.

Of the structures studied here, short period structures (i.e., those with T1 < 0.5 s) were
often observed to have different trends than those with T1 > 0.5 s in terms of how the
period range used to compute Saavg affected the dispersion of the collapse intensities.
This result is not entirely unexpected as other studies (e.g., Miranda 2001) have also
found the inelastic response of short period structures to be distinct from those of medium
and long period structures. This is related to the fact that short period systems (1) often
have periods significantly shorter than the duration of significant acceleration pulses; and
(2) undergo many more vibration cycles over a given time period than longer period
structures, and are therefore are more sensitive to cyclic deterioration and more likely to
experience ratcheting lateral deformations. The difference in the period range associated
with the smallest dispersion of a given structure was particularly evident for 4-story
generic shear walls, which had periods of T1 = 0.2 s, 0.3 s, or 0.4 s, with the smallest
ductility-dependent strength reduction factors (R = 1.5 versus 3.0 or 6.0) and the least
ductile structural components. These structures tended to have b* values significantly
smaller than the other 4-story generic shear walls.

It was also observed that the b value that produced the lowest dispersion for a given
structure was generally very dependent on the structure type. The generic MRF and RC
MRF structures had similar recommended period-independent b values (3 or 3.5 for
arithmetic spacing and 9 for logarithmic spacing) while the recommended period-
independent b value for the generic shear wall structures (6 for arithmetic spacing and 15
for logarithmic spacing) was significantly greater than that for the MRF structures. This
may be the result of fundamental differences in the way these systems respond to ground
motions (e.g., MRF systems tend to concentrate story drift demands in a few stories,
particularly for large levels of inelastic deformation, while shear wall systems tend to
have more uniform distributions of story drifts over the height of the structure) and the
Chapter 7: Average Spectral Acceleration (Saavg): Optimal Period Ranges 295

fact that the shear wall systems generally have more stories than the MRF structures for a
given fundamental period. These results suggest that different structural systems (e.g.,
braced frames and dual systems) need to be explored with regard to the effect of period
ranges on the efficiency of collapse intensities measured by Saavg before general
recommendations can be made. It should also be noted that the maximum fundamental
periods studied were T1 = 3.2 s for MRF structures and T1 = 1.6 s for generic shear wall
structures, so caution should be exercised when extrapolating these findings to longer
periods. Additionally, the generic shear wall structures used here were assumed to be
governed by flexural behavior, so the findings may not apply to shear wall structures that
fail in shear or other modes of failure (e.g., a combination of shear and flexure, out-of-
plane buckling, etc.).
Chapter 8

Summary and Conclusions

8.1 Overview
One of the most important objectives of earthquake engineering is providing protection
against structural collapse. Until recently this protection against collapse was not
explicitly quantified but instead was assumed to be sufficient for structures designed to
standards specified by building codes. Advances in computational power and the
development of models that can predict the behavior of structural components through
failure have made collapse risk assessment possible. With the advent of performance-
based earthquake engineering, which considers uncertainties in the seismic hazard and
structural response and seeks to engineer structures so that they achieve a desired level of
performance, collapse risk assessment has become increasingly important. In fact, it is a
necessary component of the performance-based earthquake engineering assessment
methodology as the probability of collapse conditioned on ground motion intensity is
required to compute expected economic losses, downtime and casualties resulting from a
seismic event.

This dissertation focuses on seismic collapse risk assessment of structures and evaluates
the effects of intensity measure (IM) selection, which is used to quantify the seismic
hazard and predict the structural response, and computational approach on the computed
risk. The main objectives of this research are:
(1) To evaluate different metrics for quantifying the collapse risk and to describe the
advantages and disadvantages of each metric.
(2) To examine and quantify the uncertainty in the collapse fragility, which describes
the probability of collapse conditioned on ground motion intensity, and in the
mean annual frequency of collapse (c) due to the number of ground motions
considered in structural response analysis.
(3) To develop efficient and reliable procedures for estimating the collapse risk.

296
Chapter 8: Summary and Conclusions 297

(4) To explain why certain IMs are able to predict structural response better (i.e.,
more efficiently) than others.
(5) To evaluate the performance of different IMs with respect to predicting the
collapse risk. Particular focus is given to the efficiency and sufficiency of the IM
and how these properties affect the computed collapse risk.
(6) To study spectral acceleration averaged over a period range (Saavg) as an IM and
provide recommendations on the period ranges that maximize the efficiency of
the IM for different structures.

8.2 Summary of findings and conclusions


The following sub-sections summarize the main conclusions of this work and are
organized by topic. Some limitations and suggestions for future work associated with
specific findings are discussed, although a more general discussion of the limitations and
future work are presented in subsequent sections.

8.2.1 Collapse risk assessment


Numerical simulation of structural collapse and the quantification of collapse risk is an
issue that has garnered significant attention from researchers in recent years. Chapter 2
presents a detailed review of the relevant research and describes two methods for
estimating collapse risk: a fully probabilistic method that does not use IMs and an IM-
based method. The latter is used throughout this work and is currently widely used due
to, among other reasons, its reduced computational effort. Different metrics used to
quantify the collapse risk were described, and the advantages and disadvantages of each
were compared. It was shown that c (or the related probability of collapse over a given
period of time) is the preferred metric because it can be used to directly compare the risks
of different structures in different sites. The reasons for this is that, unlike other collapse
risk metrics, it combines information about both the seismic hazard at the site and the
structural response at different levels of ground motion intensity, and therefore it takes
into account differences in the seismic hazard and collapse fragility at all intensities. For
these reasons, c (or the related probability of collapse over a given period of time) is
Chapter 8: Summary and Conclusions 298

used to quantify the collapse risk throughout this document. Selected simplified methods
for assessing the collapse risk of structures were also presented.

Chapter 3 presents a complete collapse risk assessment analysis using a 4-story steel
moment-resting frame (MRF) structure as a case study. The pseudo-spectral acceleration
at the first mode period of the structure, Sa(T1), is used as the IM. Note that 5%-damped
spectral ordinates are used throughout this work unless noted otherwise. In addition to
evaluating the collapse risk of the case study, the deaggregation of the collapse risk is
presented, illustrating that intensities in the lower half of the collapse fragility curve
contribute to the majority of the collapse risk for this IM. An analysis on the influence of
the shape of the seismic hazard curve, which can vary between geographic regions, on the
collapse risk is conducted. Seismic hazard curves from six locations throughout the
United States are used, and the analysis demonstrates the importance of considering
multiple intensity levels when assessing the collapse risk as the difference between the
seismic hazards of two different sites can vary significantly with the intensity level.
Furthermore, it is found that for all cases considered, despite significant differences in the
shapes of the seismic hazard curves, the most significant contribution to the collapse risk
(the peak of the collapse risk deaggregation curve) occurs at an intensity in the lower half
of the collapse fragility curve.

8.2.2 Uncertainty in the collapse fragility and mean annual frequency of


collapse
The statistical uncertainty in the collapse fragility and the mean annual frequency of
collapse (c) is addressed in Chapter 4. In particular, the uncertainty associated with the
number of ground motions used to predict the structural response is quantified, and
theoretical confidence intervals based on the binomial probability distribution are
calculated for the probability of collapse conditioned on ground motion intensity using
the Wilson score confidence intervals. The uncertainty in the collapse fragility and c as a
function of the number of ground motions used are determined for the case study
presented in Chapter 3 using bootstrap techniques, and it is demonstrated that using a
Chapter 8: Summary and Conclusions 299

small number of ground motions can produce collapse risk estimates with significant
uncertainties.

A method for estimating c by estimating the collapse fragility at only two intensity levels
was proposed. The premise of the proposed method is that for a given amount of
computational effort (as measured by the total number of nonlinear response history
analyses conducted), a better estimate of the collapse fragility can be obtained by using
more ground motions at fewer intensities compared to using more intensity levels but
fewer ground motions at each, as has typically been done in previous studies. One
advantage of the proposed method is that more attention can be paid to selection and
scaling of the ground motions to the hazard levels associated with the two intensity levels
versus an incremental dynamic analysis method that uses the same ground motions at all
intensity levels. The two intensity levels used to estimate the collapse fragility curve are
determined based on their cumulative contributions to c, as computed from a
deaggregation of c by intensity. Using an initial estimate of the collapse fragility curve,
the initial c deaggregation curve is constructed and used to identify the first intensity at
which response history analyses are conducted. The probability of collapse at that
intensity level is estimated from the results of the response history analyses, and the
collapse fragility and c deaggregation curves are updated each time a point on the
collapse fragility curve is estimated.

The ability of the proposed method to significantly reduce both the computational effort
and level of uncertainty in the c estimate was demonstrated using a case study. It was
shown that the c estimate was typically not very sensitive to the initial estimate of the
collapse fragility curve. For the case study structure, the recommended intensity levels
are those associated with a 90% and a 20% cumulative contribution to c, respectively.
Limited studies suggest that the optimum set of intensity levels may depend on the
particular collapse fragility curve and seismic hazard curve being considered, so further
investigation in this area is warranted.
Chapter 8: Summary and Conclusions 300

8.2.3 Spectral shape


Several approaches were investigated to examine the relationship between spectral shape
and nonlinear structural response, and the results are presented in Chapter 5. By
examining the equation of motion, time-domain features were identified that give rise to
significant inelastic displacements. Specifically, it was found that acceleration pulses
with significant incremental velocity (the area under the ground acceleration history
between zero-crossings) combined with movement of the system in a direction opposite
the ground acceleration during the duration of the pulse lead to large inelastic excursions.

Since a ground acceleration history can be interpreted as a series of consecutive pulses,


the response of a series of structures to individual half-sine pulses and the resulting
displacement response spectra were examined. It is shown that the displacement spectrum
of an at-rest system subjected to a half-sine acceleration pulse of arbitrary duration and
area can provide information on the pulse itself, such as the duration td and the amplitude
of the pulse. For example, a sharp increase in spectral displacement ordinates is observed
over the period range between approximately td and 3td.

Different states of the system (i.e., the displacement and velocity) at the time of the
arrival of the critical pulse, referred to as the initial conditions of the system, were
considered, and it was shown that they can significantly affect how a system responds to
a pulse, particularly when td << T, the period of the system. The largest responses
typically occur under the initial conditions that maximize the amount of time that the
system moves in the direction opposite the ground acceleration, something that is
consistent with the observations from the analysis on the equation of motion.

The peak displacement response to a ground motion was also considered and correlated
to spectral shape. Smooth spectral regions, especially smooth peaks, were found to occur
when a single segment of the ground acceleration controls the peak response of a period
range. Sharp changes in spectral shape, such as a V-shaped valley, typically occur
when there is a change in the controlling ground acceleration history segment. Both
elastic and inelastic systems were considered, and it was shown that their peak responses
Chapter 8: Summary and Conclusions 301

and spectral shapes can differ significantly. This is primarily attributed to the different
initial conditions at the beginning of a given ground motion segment in the inelastic
system relative to the elastic system caused by a previous yielding in the system.

The spectral shape of damaging records was also examined using results from the case
study of Chapter 3. Damaging records are defined as those with low collapse intensities
Sa(T1)col. It was found that the spectral shape of damaging records is, generally, distinct
from that of more benign records. Specifically, damaging records typically exhibit
sharply increasing spectral displacements for T > T1. The examination of selected
damaging records as well as displacement histories of the case study structure yielded
analogous results to previous approaches, namely that (1) the shape of the spectrum can
provide information about underlying pulses in the ground motion; and (2) pulses with
significant incremental velocities can cause large displacement excursions, particularly if
the system moves in the same direction as the momentum imparted by the pulse.

The relationship between spectral shape and the collapse intensity of the record was also
discussed, and three measures of spectral shape were evaluated with respect to their
ability to predict the collapse intensity. These measures included a direct measure of
spectral shape, called SaRatio, and indirect measures of spectral shape and , which
measure the difference between the spectral ordinate or peak ground velocity (PGV) of
the ground motion and the mean prediction from a ground motion prediction equation
(GMPE). SaRatio, which is defined as the ratio of Sa(T1) to the geometric mean of
pseudo-spectral acceleration values over a certain period range (Saavg), performed
significantly better than the other spectral shape parameters. This was attributed to the
fact that it directly measures the spectrum of a record over a range of periods, as opposed
to indirect spectral shape measures that only consider a single spectral ordinate (or a
linear combination of an ordinate and PGV in the case of ) and measure the ordinate(s)
with reference to a GMPE (which provides the average ordinate based on many records).

Since seismic hazard deaggregations do not currently provide information about SaRatio,
a method for estimating a target SaRatio value was proposed. In this method, SaRatio is
Chapter 8: Summary and Conclusions 302

computed from the spectrum of an event dominating the seismic hazard or a mean causal
event. Two approximate methods that account for the effect of SaRatio in collapse risk
assessment were discussed. One is appropriate for use with the results from incremental
dynamic analyses while the other is appropriate when estimating the probability of
collapse conditioned on the ground motion intensity.

8.2.4 Intensity measures for predicting collapse


Chapter 6 evaluates the performance of different IMs in predicting the collapse risk of a
structure. The IMs are evaluated with respect to desirable properties of an IM including
efficiency, sufficiency, the robustness of the collapse risk estimate, and the ability to
compute hazard information. The geometric mean of the pseudo-spectral acceleration
over a period range (Saavg) was shown to be consistently and significantly more efficient
than Sa(T1) at predicting the collapse intensity of a ground motion. Case studies of nearly
700 moment-resisting frame and shear wall structures showed that Saavg reduced the
dispersion in the collapse intensities by over 40%, on average, compared to Sa(T1). The
sufficiency of Saavg with respect to magnitude, distance and spectral shape measures was
evaluated using the same case studies, and Saavg was generally shown to be sufficient
with respect to these parameters.

The generally superior performance of Saavg compared to Sa(T1) was attributed to that
fact that Saavg accounts directly for spectral shape features and indirectly for the
underlying pulses or ground motion segments that cause damage, whereas Sa(T1)
measures the elastic response at only a single period, which can be heavily influenced by
the initial conditions at the onset of a critical ground motion pulse. Stated somewhat
differently, the response spectrum can be viewed as a signature of the duration and
intensity of individual pulses or pulse segments within the ground acceleration combined
with the initial conditions/state of the system on the onset of the pulse. Saavg
incorporates more information about the ground motion by taking the average value of
the signature over several points as opposed to Sa(T1), which takes the value of the
signature at a single period and therefore can be strongly affected by the state of the
Chapter 8: Summary and Conclusions 303

system at the onset of the pulse. The additional information about the record that Saavg
incorporates is what makes it a more efficient measure of collapse intensity than Sa(T1).

Obtaining a hazard curve for Saavg based on existing GMPEs that predict spectral
acceleration values was discussed. The performance of different IMs in predicting the
collapse risk was evaluated and compared using a case study structure and seven different
ground motion sets. Sa(T1), Saavg, and PGV were the IMs considered. Compared to Saavg,
both Sa(T1) and PGV generally produced higher and significantly more varied (i.e., less
robust) collapse risk estimates among the different ground motion sets. The higher
collapse risk estimates were attributed primarily to the decreased efficiency of these IMs
(i.e., the increased dispersion in structural response) compared to Saavg and, in the case of
Sa(T1), the increased dispersion in the value predicted by the GMPE. That is, compared
to Sa(T1) and PGV, Saavg was found to be predicted with less or comparable dispersion
from a GMPE perspective, to be a more efficient and reliable predictor of collapse, and to
produce collapse risk estimates that were significantly more robust (less variable) with
respect to the particular ground motions used to compute the structural response, even in
cases where the collapse risk estimates computed using Sa(T1) accounted for the spectral
shape of the records used. Though it may be theoretically possible to compute the same
(correct) collapse risk using any IM (see, e.g., Bradley 2012; Lin 2012; Lin et al.
2013ba), in practice collapse risk assessment involves using a limited number of ground
motions to estimate the structural response and, in some cases, there may be a lack of
available ground motions that are compatible with the ground motions expected at a
particular site. Because of these practical issues, there is a benefit to using an IM that is
an efficient predictor of structural response and sufficient with respect to ground motion
properties when assessing the collapse risk.

The relatively stable (i.e., robust) collapse risk estimates obtained across the ground
motion sets when using Saavg is an indicator of the sufficiency of this IM with respect to
various ground motion properties and suggests that collapse risk estimates are not
particularly sensitive to the ground motions used in structural response prediction when
using this IM. This implies that the careful record selection and/or modification of
Chapter 8: Summary and Conclusions 304

structural response results required to obtain a good estimate of the collapse risk when
using Sa(T1) as the IM may not be required if Saavg is the IM. Additional studies are
needed to confirm this, however, as the results presented here are based on a single case
study.

8.2.5 Effect of period range on the efficiency of Saavg


Chapter 7 investigates how the period range used to compute Saavg affects the efficiency
of structural response predictions as measured by the dispersion of the collapse
intensities. Nearly 700 structures including SDOF systems and generic MRF, generic
shear wall, and reinforced concrete (RC) MRF systems are used to identify trends
between the optimum period range (i.e., that with the highest efficiency) and structural
properties such as T1 and structural system. Structures with fundamental periods between
T1 = 0.1 s and 3.2 s and between 1 and 20 stories were considered. Let Saavg(a, b) denote
the value of Saavg computed using a period range between aT1 and bT1, where a and b
are non-negative constants such that a b. Let a* and b* denote the values of a and b,
respectively, that are associated with the overall minimum dispersion computed for a
given structure using a given set of ground motions.

It was observed that the period ranges that tended to produce the smallest dispersions in
collapse intensities were primarily a function of the structural system (i.e., MRF versus
shear wall) and of the fundamental period of the system. With the exception of short
period structures (i.e., those with T1 < 0.5 s), it was observed that the value of b*
generally decreased as the fundamental period of the structure increased. The average
value of the minimum dispersion computed for a given structure was approximately
lnSaAvg(a*,b*) = 0.20 for the generic MRF structures and lnSaAvg(a*,b*) = 0.16 for both the RC
MRF and the generic shear wall structures. Compared to the dispersion computed using
Sa(T1), the dispersions computed using Saavg(a*, b*) represent average reductions of
nearly 50% for the generic MRF structures and over 60% for the RC MRF and generic
shear wall structures. It was observed that the spacing of periods (i.e., arithmetic versus
logarithmic) within the period range used to compute Saavg generally did not affect the
minimum dispersion of a given structure; however, the period range producing the
Chapter 8: Summary and Conclusions 305

minimum dispersion was typically very different for each of the two period spacing
schemes.

When the periods used to compute Saavg are spaced arithmetically within the period
range, it was observed that the dispersion of collapse intensities was generally much more
sensitive to the b value than the a value. It was shown that using Saavg(0, 3) for the MRF
structures (i.e., using a period range between 0 and 3T1) and Saavg(0, 6) for the shear wall
structures (i.e., using a period range between 0 and 6T1) generally produced dispersions
that were close to the dispersion computed using Saavg(a*, b*) (i.e., the overall minimum
dispersion), with an average difference in dispersion less than 0.04 lnSaAvg units.

When the periods used to compute Saavg are spaced logarithmically within the period
range, it was observed that the dispersion of collapse intensities was sensitive to both the
a and the b values, particularly if the value of a and/or b was less than the value of a*
and/or b*, respectively. It was shown that using Saavg(0.2, 9) for the generic MRF
structures, Saavg(0.4, 9) for the RC MRF structures, and Saavg(0.7, 15) for the shear wall
structures generally produced dispersions that were close to the dispersion computed
using Saavg(a*, b*), with an average difference in dispersion less than 0.04 lnSaAvg units.
It should be noted that periods longer than 10 s were not used to compute Saavg even if
bT1 exceeded 10 s. This limit was imposed because of limitation associated with the
maximum usable period of recordings and the lack of GMPEs predicting spectral
ordinates at periods greater than 10 s.

When the period range (i.e., the a and b values) used to compute Saavg was determined
based on regressions of a* versus T1 and b* versus T1 (which were specific to the
structural system and period spacing scheme used), it was observed that a slight reduction
in the dispersion was achieved with respect to the dispersion calculated using the period-
independent values of a and b (e.g., using Saavg(0, 6) with arithmetically spaced periods
for the shear wall structures). The average difference in the dispersions produced by the
period-dependent versus the period-independent values of a and b was less than 0.03
Chapter 8: Summary and Conclusions 306

lnSaAvg units. Given this relatively small difference it is recommended to use the period-
independent values of a and b when computing Saavg.

The values of a and b that led to the minimum dispersion in the collapse intensities were
significantly higher for logarithmic spacing than for arithmetic spacing. This results from
the fact that logarithmic spacing samples more heavily from the lower half of the period
range. The increase in the a value for logarithmic spacing is thought to be the result of the
fact that for medium to long period structures, the parts of the ground motion that control
the response of elastic SDOF systems (and produce the short-period spectral ordinates)
are generally not the same as those controlling the response of longer period systems.
Thus, there is a need to avoid over sampling in this period region, which is why the value
of a increases from the value of 0 that produced small dispersions when using arithmetic
period spacing to values of at least 0.2 for logarithmic period spacing. The increase in the
b value may be the related to the increase in a, as increasing the b value for a given a
value will lessen the concentration of samples at shorter periods.

It is observed that when using the previously recommended period-independent values of


a and b (which depend on the period spacing scheme and structure class) to compute
Saavg, logarithmic spacing generally results in a slightly decreased dispersion in the
collapse intensities compared to arithmetic period spacing, with an average difference of
0.02 lnSaAvg units or less. Given this small difference, one might prefer to use arithmetic
period spacing because the recommended b value for a given structural system is
generally much lower for arithmetic spacing versus logarithmic spacing. Using a lower b
value reduces potential issues associated with a lack of seismic hazard data at long
periods, the maximum usable period of a record, and record processing effects that may
affect the amplitude of spectral ordinates at long periods.

Regarding the different period ranges recommended for different structural systems, this
may be the result of fundamental differences in the way these systems respond to ground
motions (e.g., MRF systems tend to concentrate story drift demands in a few stories,
particularly for large levels of inelastic deformation, while shear wall systems tend to
Chapter 8: Summary and Conclusions 307

have more uniform distributions of story drifts along the height of the building) and the
fact that for the structures considered the shear wall systems generally have more stories
than the MRF systems. These results suggest that different structural systems (e.g.,
braced frames or dual systems) need to be explored with regard to the effect of period
ranges on the efficiency of collapse intensities measured by Saavg before general
recommendations can be made. All results discussed here were based on computing Saavg
using 100 points/periods within a given period range. The sensitivity of the results to the
number of periods was studied, and it was found that the results were not significantly
affected when only 10 points/periods were used to compute Saavg.

8.3 Limitations and suggestions for future work


A significant limitation of this work is that it is focused on collapse and does not directly
consider other levels or descriptions of structural response. The findings regarding the
effect of spectral shape on structural response, the efficiency and sufficiency of Saavg as
well as the robustness of risk estimates based on this IM and the recommended period
ranges for computing Saavg are all specific to collapse and may not apply to other
descriptions of structural response such as the displacement and acceleration responses
when the structure does not collapse. However, as the sidesway collapse mode considered
in this work is essentially an excessive displacement response, it is anticipated that many
of the findings will be applicable to displacement-based responses, especially for large
levels of nonlinearity; however, future work is required to confirm this.

The proposed method for estimating c presented in Section 4.5 has been demonstrated
for a case study, but more studies are needed to validate its ability to reduce the
uncertainty and/or computational effort in the c estimate. Additionally, the intensity
levels recommended for estimating the collapse fragility are specific to the case study
(i.e., the particular structure and seismic hazard curve), although limited results suggest
that intensities with cumulative contributions to c of at least 70% for the first intensity
and less than or equal to 35% for the second intensity should be used. Future work on the
optimum intensity levels to use for estimating c based on studies of a variety of
structures and seismic hazard curves will be useful.
Chapter 8: Summary and Conclusions 308

The previous recommendations for the intensity levels to use when estimating the
collapse fragility are also conditional on the IM being Sa(T1). As indicated by the c
deaggregation presented in Figure 6.10, the collapse risk may not be dominated by
intensities in the lower tail of the collapse fragility curve when Saavg is the IM, which
could affect the intensity levels that are best for estimating the collapse risk. The
proposed method for estimating c is certainly applicable when using Saavg as the IM, and
in fact using Saavg instead of Sa(T1) is expected to provide a better estimate of c as Saavg
was demonstrated to generally be more efficient, sufficient, and produce more robust
structural response estimates than Sa(T1). However, the optimum intensity levels to use
with Saavg were not explored, and future work in this area is necessary.

In Chapter 5 it is shown that there is a strong relationship between the spectral shape as
measured by SaRatio and the collapse intensity Sa(T1)col and that the SaRatios of the
records used to predict the structural response can affect the computed collapse risk. The
approximate methods discussed in Section 5.7.2 for incorporating SaRatio in collapse
risk assessment should be used with caution. The approximate methods for estimating a
target SaRatio are reasonable approaches based on available data regarding the expected
spectral shape; however, future research on obtaining an appropriate target SaRatio is
necessary, specifically knowledge of the distribution of SaRatio given Sa(T1) obtained
from a probabilistic seismic hazard analysis (PSHA) deaggregation. A more direct way to
incorporate the effect of spectral shape when using Sa(T1) as the IM would be to use
records with appropriate spectral shapes in terms of expected values and variability, such
as selecting records to match a conditional spectrum based on a PSHA deaggregation
(see, e.g., Lin et al. 2013ba).

In Chapter 6 the GMPE for Saavg, which was used to compute the seismic hazard curve,
was obtained using GMPEs for Sa(T) and equations for the correlation of spectral values
at different periods because an established GMPE based directly on Saavg does not exist.
Though the former method is convenient because it uses existing equations, it will be
useful to confirm that it produces comparable values to a GMPE derived from regressions
directly on the Saavg values of records. As explained in Section 6.3.4, potential
Chapter 8: Summary and Conclusions 309

differences in values could arise due to number of factors including the fact the equations
of the GMPE for Sa(T) and of the correlations between spectral values at different
periods are not exact matches to the empirical data. The impact of any differences should
be quantified, especially differences in the resulting seismic hazard curve and collapse
risk estimate.

On a related note, a significant obstacle to the use of Saavg in engineering practice is that
hazard curves and PSHA deaggregations are not widely available for Saavg as they are for
Sa(T1). The work presented in Chapter 7 suggests that the optimum period range to use
for Saavg varies with the structural system, so hazard curves for Saavg computed over
many different period ranges will be desirable to capture both the period range that works
for a particular structural system (e.g., 0 to 3T1) and the range of fundamental periods of
individual structures. Although this will increase the computational effort, it may not be
insurmountable, particularly if the GMPE for Saavg is based on existing GMPEs for Sa(T)
and equations for the correlation coefficients between spectral values at different periods
versus developing GMPEs based on regressions of the Saavg values of individual records.
It is envisioned that seismic hazard analysis applications such as those provided by
OpenSHA (Field et al. 2003; Field 2013) and the USGS (USGS 2012b; USGS 2012a)
can incorporate features allowing the user to input the specific periods he/she wishes to
use to compute Saavg.

The conclusions drawn about the efficiency, sufficiency, and recommended period range
for Saavg in Chapters 6 and 7 are obviously limited by the case studies considered. For
example, only structures with fundamental periods between T1 = 0.1 s and 3.2 s were
investigated, so longer periods structures are not considered in this study. Additionally,
the generic shear wall structures were assumed to fail in flexure, so the findings may not
apply to shear walls that fail in shear or other modes of failure (e.g., a combination of
shear and flexure, out-of-plane buckling, etc.). Only shear walls and MRF structural
systems were analyzed, so future investigations could examine the performance of Saavg
for other structural systems, including braced frames, base-isolated structures, and
structures with dual lateral systems. This is particularly important because it was found
Chapter 8: Summary and Conclusions 310

that the recommended period range for computing Saavg was different for the MRF versus
the shear wall structures. Evaluation of Saavg with respect to specific types of ground
motions sets (e.g., those comprised exclusively of records with near-fault, forward
directivity pulses or long durations records) were not explicitly considered, so future
research in this area may be necessary if this is of concern.

8.4 Implications for engineering practice


In design codes such as ASCE 7-10 (ASCE 2010), ground motions used in response
history analysis must be scaled so that the average response spectrum of the ground
motion set is not less than the design response spectrum at periods between 0.2T1 and
1.5T1. Results presented in Chapters 5, 6 and 7 suggest that regions of the spectrum up to
3T1 for MRFs and up to 6T1 for shear walls are important for collapse, as including
periods up to these values when computing the average spectral value improves
prediction of structural response (either via the spectral shape parameter SaRatio or when
using Saavg as an IM). At the design level, the structure should experience significantly
smaller levels of nonlinearity than those associated with collapse, so future work will be
valuable in determining whether there is a benefit to extending the period range over
which the spectra are compared past 1.5T1 for design level checks (i.e., whether regions
of the spectrum greater than 1.5T1 are important for the response of the structure) and, if
so, determining how far to extend the period range. Limited results based on the case
study of the MRF structure presented in Chapter 3 show that prediction of the maximum
interstory drift ratio response at the 10/50 hazard level is improved when the upper limit
of the period range used to compute the average spectral value is increased from 1.5T1 to
a value between 2T1 and 3T1. As previously stated, however, comprehensive studies are
required to determine whether the period range used with respect to the scaling of ground
motions should be extended past 1.5T1.

8.5 Concluding remarks


This work contributes to improving the seismic collapse risk assessment of buildings by
examining the effects of collapse risk computation (in terms of the collapse risk metric
chosen as well as the intensity levels and number of ground motions used in structural
Chapter 8: Summary and Conclusions 311

response assessment) and the intensity measure used to quantify the ground motion
intensity and predict the structural response. The findings of this work should be
interpreted considering the limitations of the studies. The promising results presented
here regarding the use of spectral acceleration averaged over a period range demonstrate
that it can be a particularly efficient predictor of structural response, particularly collapse,
and when used as an IM can produce significantly more reliable and robust collapse risk
estimates compared to other IMs. However, additional case studies regarding the collapse
risk estimates obtained using Saavg are required to validate these findings. Future work
can also expand recommendations on the period range used to compute Saavg to
additional structures.
Appendix A

MRCD 137 Ground Motion Set

This appendix documents the ground motions used in the case study. All ground motions
are from the PEER Next Generation Attenuation (NGA) database (Chiou et al. 2008).
This ground motion set, denoted as MRCD 137, is comprised of all NGA records,
excluding those from dam abutments, that are within a specified range of magnitude (Mw
between 6.93 and 7.62), Joyner-Boore distance (Rjb between 0 and 27 km), NEHRP site
class (C or D), and are from strike-slip, reverse, or reverse-oblique faults. The ground
motion set consists of 137 acceleration records (each with two horizontal components)
from 11 different events. Approximately 60% of the records are from the Mw = 7.62,
1999 Chi-Chi, Taiwan event, which is reflected in the scatter plot in Figure A.1 that
shows the distribution of Mw and Rjb for the 137 records. The pseudo-acceleration
response spectra of the 137 records are shown in Figure A.2, where the spectral ordinates
are the 5%-damped, orientation-independent measures reported by the NGA (Chiou et al.
2008; Boore et al. 2006). A summary of the MRCD 137 ground motion set including
event data and number of recordings per event is provided in Table A.1, and details about
each of the 137 records including NGA identification numbers and station names are
provided in Table A.2.

312
Appendix A: MRCD 137 Ground Motion Set 313

Figure A.1. Magnitudes and distances of records in the MRCD 137 set.

Figure A.2. Response spectra of records in the MRCD 137 set with the median spectrum
denoted by the thick black line.
Appendix A: MRCD 137 Ground Motion Set 314

Table A.1. Summary of MRCD 137 set


No. of Records
Year Event Mw
in Set
1940 Imperial Valley 6.95 1
1978 Tabas, Iran 7.35 2
1989 Loma Prieta 6.93 22
1992 Cape Mendocino 7.01 5
1992 Landers 7.28 8
1999 Kocaeli, Turkey 7.51 3
1999 Chi-Chi, Taiwan 7.62 84
1999 Duzce, Turkey 7.14 9
1979 St Elias, Alaska 7.54 1
1990 Manjil, Iran 7.37 1
1999 Hector Mine 7.13 1

Table A.2. Records comprising the MRCD 137 ground motion set
Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault ClstD -Boore Vs30
No. Rec. Year Station Name
(Mw) Name Type (km) Dist. (m/s)
No.
(km)
Imperial Strike- El Centro Array
1 6 6.95 1940 6.1 6.1 213
Valley-02 Slip #9
2 138 7.35 Tabas, Iran 1978 Reverse Boshrooyeh 28.8 24.1 339
3 139 7.35 Tabas, Iran 1978 Reverse Dayhook 13.9 0.0 660
Reverse- Agnews State
4 737 6.93 Loma Prieta 1989 24.6 24.3 240
Oblique Hospital
Reverse- Anderson Dam
5 739 6.93 Loma Prieta 1989 20.3 19.9 489
Oblique (Downstream)
Reverse-
6 741 6.93 Loma Prieta 1989 BRAN 10.7 3.9 376
Oblique
Reverse-
7 752 6.93 Loma Prieta 1989 Capitola 15.2 8.7 289
Oblique
Reverse-
8 753 6.93 Loma Prieta 1989 Corralitos 3.9 0.2 462
Oblique
Reverse- Coyote Lake
9 754 6.93 Loma Prieta 1989 20.8 20.4 295
Oblique Dam (Downst)
Reverse- Gilroy -
10 763 6.93 Loma Prieta 1989 10.0 9.2 730
Oblique Gavilan Coll.
Reverse- Gilroy -
11 764 6.93 Loma Prieta 1989 11.0 10.3 339
Oblique Historic Bldg.
Reverse-
12 766 6.93 Loma Prieta 1989 Gilroy Array #2 11.1 10.4 271
Oblique
Appendix A: MRCD 137 Ground Motion Set 315

Table A.2 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault ClstD -Boore Vs30
No. Rec. Year Station Name
(Mw) Name Type (km) Dist. (m/s)
No.
(km)
Reverse-
13 767 6.93 Loma Prieta 1989 Gilroy Array #3 12.8 12.2 350
Oblique
Reverse-
14 768 6.93 Loma Prieta 1989 Gilroy Array #4 14.3 13.8 222
Oblique
Reverse-
15 769 6.93 Loma Prieta 1989 Gilroy Array #6 18.3 17.9 663
Oblique
Reverse-
16 770 6.93 Loma Prieta 1989 Gilroy Array #7 22.7 22.4 334
Oblique
Reverse- Hollister Diff.
17 778 6.93 Loma Prieta 1989 24.8 24.5 216
Oblique Array
Reverse-
18 779 6.93 Loma Prieta 1989 LGPC 3.9 0.0 478
Oblique
San Jose -
Reverse-
19 801 6.93 Loma Prieta 1989 Santa Teresa 14.7 14.2 672
Oblique
Hills
Reverse- Saratoga -
20 802 6.93 Loma Prieta 1989 8.5 7.6 371
Oblique Aloha Ave
Reverse- Saratoga - W
21 803 6.93 Loma Prieta 1989 9.3 8.5 371
Oblique Valley Coll.
Reverse- Sunnyvale -
22 806 6.93 Loma Prieta 1989 24.2 23.9 268
Oblique Colton Ave.
Reverse-
23 809 6.93 Loma Prieta 1989 UCSC 18.5 12.2 714
Oblique
Reverse- UCSC Lick
24 810 6.93 Loma Prieta 1989 18.4 12.0 714
Oblique Observatory
Reverse-
25 811 6.93 Loma Prieta 1989 WAHO 17.5 11.0 376
Oblique
Cape Cape
26 825 7.01 1992 Reverse 7.0 0.0 514
Mendocino Mendocino
Cape Fortuna -
27 827 7.01 1992 Reverse 20.0 16.0 457
Mendocino Fortuna Blvd
Cape
28 828 7.01 1992 Reverse Petrolia 8.2 0.0 713
Mendocino
Cape Rio Dell
29 829 7.01 1992 Reverse 14.3 7.9 312
Mendocino Overpass - FF
Cape Shelter Cove
30 830 7.01 1992 Reverse 28.8 26.5 514
Mendocino Airport
Strike-
31 848 7.28 Landers 1992 Coolwater 19.7 19.7 271
Slip
Strike- Desert Hot
32 850 7.28 Landers 1992 21.8 21.8 345
Slip Springs
Appendix A: MRCD 137 Ground Motion Set 316

Table A.2 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault ClstD -Boore Vs30
No. Rec. Year Station Name
(Mw) Name Type (km) Dist. (m/s)
No.
(km)
Strike-
33 864 7.28 Landers 1992 Joshua Tree 11.0 11.0 379
Slip
Strike-
34 879 7.28 Landers 1992 Lucerne 2.2 2.2 685
Slip
Strike- Mission Creek
35 880 7.28 Landers 1992 27.0 27.0 345
Slip Fault
Strike- Morongo
36 881 7.28 Landers 1992 17.3 17.3 345
Slip Valley
Strike- North Palm
37 882 7.28 Landers 1992 26.8 26.8 345
Slip Springs
Strike- Yermo Fire
38 900 7.28 Landers 1992 23.6 23.6 354
Slip Station
Kocaeli, Strike-
39 1148 7.51 1999 Arcelik 13.5 10.6 523
Turkey Slip
Kocaeli, Strike-
40 1158 7.51 1999 Duzce 15.4 13.6 276
Turkey Slip
Kocaeli, Strike-
41 1176 7.51 1999 Yarimca 4.8 1.4 297
Turkey Slip
Chi-Chi, Reverse-
42 1178 7.62 1999 ALS 10.8 0.7 553
Taiwan Oblique
Chi-Chi, Reverse-
43 1180 7.62 1999 CHY002 25.0 25.0 235
Taiwan Oblique
Chi-Chi, Reverse-
44 1182 7.62 1999 CHY006 9.8 9.8 438
Taiwan Oblique
Chi-Chi, Reverse-
45 1184 7.62 1999 CHY010 20.0 19.9 474
Taiwan Oblique
Chi-Chi, Reverse-
46 1193 7.62 1999 CHY024 9.6 9.6 428
Taiwan Oblique
Chi-Chi, Reverse-
47 1194 7.62 1999 CHY025 19.1 19.1 278
Taiwan Oblique
Chi-Chi, Reverse-
48 1197 7.62 1999 CHY028 3.1 3.1 543
Taiwan Oblique
Chi-Chi, Reverse-
49 1198 7.62 1999 CHY029 11.0 11.0 545
Taiwan Oblique
Chi-Chi, Reverse-
50 1201 7.62 1999 CHY034 14.8 14.8 379
Taiwan Oblique
Chi-Chi, Reverse-
51 1202 7.62 1999 CHY035 12.7 12.6 474
Taiwan Oblique
Chi-Chi, Reverse-
52 1203 7.62 1999 CHY036 16.1 16.1 233
Taiwan Oblique
Appendix A: MRCD 137 Ground Motion Set 317

Table A.2 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault ClstD -Boore Vs30
No. Rec. Year Station Name
(Mw) Name Type (km) Dist. (m/s)
No.
(km)
Chi-Chi, Reverse-
53 1205 7.62 1999 CHY041 19.8 19.4 492
Taiwan Oblique
Chi-Chi, Reverse-
54 1208 7.62 1999 CHY046 24.1 24.1 442
Taiwan Oblique
Chi-Chi, Reverse-
55 1209 7.62 1999 CHY047 24.1 24.1 273
Taiwan Oblique
Chi-Chi, Reverse-
56 1227 7.62 1999 CHY074 10.8 0.7 553
Taiwan Oblique
Chi-Chi, Reverse-
57 1231 7.62 1999 CHY080 2.7 0.1 553
Taiwan Oblique
Chi-Chi, Reverse-
58 1238 7.62 1999 CHY092 22.7 22.7 254
Taiwan Oblique
Chi-Chi, Reverse-
59 1244 7.62 1999 CHY101 10.0 10.0 259
Taiwan Oblique
Chi-Chi, Reverse-
60 1246 7.62 1999 CHY104 18.0 18.0 223
Taiwan Oblique
Chi-Chi, Reverse-
61 1403 7.62 1999 NSY 13.2 13.2 600
Taiwan Oblique
Chi-Chi, Reverse-
62 1462 7.62 1999 TCU 5.2 5.2 473
Taiwan Oblique
Chi-Chi, Reverse-
63 1480 7.62 1999 TCU036 19.8 19.8 273
Taiwan Oblique
Chi-Chi, Reverse-
64 1481 7.62 1999 TCU038 25.4 25.4 273
Taiwan Oblique
Chi-Chi, Reverse-
65 1482 7.62 1999 TCU039 19.9 19.9 541
Taiwan Oblique
Chi-Chi, Reverse-
66 1483 7.62 1999 TCU040 22.1 22.1 362
Taiwan Oblique
Chi-Chi, Reverse-
67 1484 7.62 1999 TCU042 26.3 26.3 273
Taiwan Oblique
Chi-Chi, Reverse-
68 1485 7.62 1999 TCU045 26.0 26.0 705
Taiwan Oblique
Chi-Chi, Reverse-
69 1486 7.62 1999 TCU046 16.7 16.7 466
Taiwan Oblique
Chi-Chi, Reverse-
70 1488 7.62 1999 TCU048 13.6 13.6 551
Taiwan Oblique
Chi-Chi, Reverse-
71 1489 7.62 1999 TCU049 3.8 3.8 487
Taiwan Oblique
Chi-Chi, Reverse-
72 1490 7.62 1999 TCU050 9.5 9.5 273
Taiwan Oblique
Chi-Chi, Reverse-
73 1491 7.62 1999 TCU051 7.7 7.7 273
Taiwan Oblique
Appendix A: MRCD 137 Ground Motion Set 318

Table A.2 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault ClstD -Boore Vs30
No. Rec. Year Station Name
(Mw) Name Type (km) Dist. (m/s)
No.
(km)
Chi-Chi, Reverse-
74 1492 7.62 1999 TCU052 0.7 0.0 579
Taiwan Oblique
Chi-Chi, Reverse-
75 1493 7.62 1999 TCU053 6.0 6.0 455
Taiwan Oblique
Chi-Chi, Reverse-
76 1494 7.62 1999 TCU054 5.3 5.3 461
Taiwan Oblique
Chi-Chi, Reverse-
77 1495 7.62 1999 TCU055 6.4 6.4 273
Taiwan Oblique
Chi-Chi, Reverse-
78 1496 7.62 1999 TCU056 10.5 10.5 273
Taiwan Oblique
Chi-Chi, Reverse-
79 1497 7.62 1999 TCU057 11.8 11.8 474
Taiwan Oblique
Chi-Chi, Reverse-
80 1498 7.62 1999 TCU059 17.1 17.1 273
Taiwan Oblique
Chi-Chi, Reverse-
81 1499 7.62 1999 TCU060 8.5 8.5 273
Taiwan Oblique
Chi-Chi, Reverse-
82 1500 7.62 1999 TCU061 17.2 17.2 273
Taiwan Oblique
Chi-Chi, Reverse-
83 1501 7.62 1999 TCU063 9.8 9.8 476
Taiwan Oblique
Chi-Chi, Reverse-
84 1502 7.62 1999 TCU064 16.6 16.6 273
Taiwan Oblique
Chi-Chi, Reverse-
85 1503 7.62 1999 TCU065 0.6 0.6 306
Taiwan Oblique
Chi-Chi, Reverse-
86 1504 7.62 1999 TCU067 0.6 0.6 434
Taiwan Oblique
Chi-Chi, Reverse-
87 1505 7.62 1999 TCU068 0.3 0.0 487
Taiwan Oblique
Chi-Chi, Reverse-
88 1506 7.62 1999 TCU070 19.0 19.0 401
Taiwan Oblique
Chi-Chi, Reverse-
89 1507 7.62 1999 TCU071 5.3 0.0 625
Taiwan Oblique
Chi-Chi, Reverse-
90 1508 7.62 1999 TCU072 7.0 0.0 468
Taiwan Oblique
Chi-Chi, Reverse-
91 1509 7.62 1999 TCU074 13.5 0.0 549
Taiwan Oblique
Chi-Chi, Reverse-
92 1510 7.62 1999 TCU075 0.9 0.9 573
Taiwan Oblique
Chi-Chi, Reverse-
93 1511 7.62 1999 TCU076 2.8 2.8 615
Taiwan Oblique
Chi-Chi, Reverse-
94 1512 7.62 1999 TCU078 8.2 0.0 443
Taiwan Oblique
Appendix A: MRCD 137 Ground Motion Set 319

Table A.2 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault ClstD -Boore Vs30
No. Rec. Year Station Name
(Mw) Name Type (km) Dist. (m/s)
No.
(km)
Chi-Chi, Reverse-
95 1513 7.62 1999 TCU079 11.0 0.0 364
Taiwan Oblique
Chi-Chi, Reverse-
96 1515 7.62 1999 TCU082 5.2 5.2 473
Taiwan Oblique
Chi-Chi, Reverse-
97 1517 7.62 1999 TCU084 11.2 0.0 553
Taiwan Oblique
Chi-Chi, Reverse-
98 1519 7.62 1999 TCU087 7.0 7.0 474
Taiwan Oblique
Chi-Chi, Reverse-
99 1520 7.62 1999 TCU088 18.2 4.7 553
Taiwan Oblique
Chi-Chi, Reverse-
100 1521 7.62 1999 TCU089 8.9 0.0 553
Taiwan Oblique
Chi-Chi, Reverse-
101 1527 7.62 1999 TCU100 11.4 11.4 474
Taiwan Oblique
Chi-Chi, Reverse-
102 1528 7.62 1999 TCU101 2.1 2.1 273
Taiwan Oblique
Chi-Chi, Reverse-
103 1529 7.62 1999 TCU102 1.5 1.5 714
Taiwan Oblique
Chi-Chi, Reverse-
104 1530 7.62 1999 TCU103 6.1 6.1 494
Taiwan Oblique
Chi-Chi, Reverse-
105 1531 7.62 1999 TCU104 12.9 12.9 474
Taiwan Oblique
Chi-Chi, Reverse-
106 1532 7.62 1999 TCU105 17.2 17.2 576
Taiwan Oblique
Chi-Chi, Reverse-
107 1533 7.62 1999 TCU106 15.0 15.0 474
Taiwan Oblique
Chi-Chi, Reverse-
108 1534 7.62 1999 TCU107 16.0 16.0 474
Taiwan Oblique
Chi-Chi, Reverse-
109 1535 7.62 1999 TCU109 13.1 13.1 474
Taiwan Oblique
Chi-Chi, Reverse-
110 1536 7.62 1999 TCU110 11.6 11.6 213
Taiwan Oblique
Chi-Chi, Reverse-
111 1537 7.62 1999 TCU111 22.1 22.1 238
Taiwan Oblique
Chi-Chi, Reverse-
112 1540 7.62 1999 TCU115 21.8 21.8 215
Taiwan Oblique
Chi-Chi, Reverse-
113 1541 7.62 1999 TCU116 12.4 12.4 493
Taiwan Oblique
Chi-Chi, Reverse-
114 1542 7.62 1999 TCU117 25.4 25.4 199
Taiwan Oblique
Chi-Chi, Reverse-
115 1543 7.62 1999 TCU118 26.8 26.8 215
Taiwan Oblique
Appendix A: MRCD 137 Ground Motion Set 320

Table A.2 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault ClstD -Boore Vs30
No. Rec. Year Station Name
(Mw) Name Type (km) Dist. (m/s)
No.
(km)
Chi-Chi, Reverse-
116 1545 7.62 1999 TCU120 7.4 7.4 459
Taiwan Oblique
Chi-Chi, Reverse-
117 1546 7.62 1999 TCU122 9.4 9.4 475
Taiwan Oblique
Chi-Chi, Reverse-
118 1547 7.62 1999 TCU123 14.9 14.9 273
Taiwan Oblique
Chi-Chi, Reverse-
119 1548 7.62 1999 TCU128 13.2 13.2 600
Taiwan Oblique
Chi-Chi, Reverse-
120 1549 7.62 1999 TCU129 1.8 1.8 664
Taiwan Oblique
Chi-Chi, Reverse-
121 1550 7.62 1999 TCU136 8.3 8.3 474
Taiwan Oblique
Chi-Chi, Reverse-
122 1551 7.62 1999 TCU138 9.8 9.8 653
Taiwan Oblique
Chi-Chi, Reverse-
123 1553 7.62 1999 TCU141 24.2 24.2 215
Taiwan Oblique
Chi-Chi, Reverse-
124 1595 7.62 1999 WGK 10.0 10.0 259
Taiwan Oblique
Chi-Chi, Reverse-
125 1596 7.62 1999 WNT 1.8 1.8 664
Taiwan Oblique
Duzce, Strike-
126 1602 7.14 1999 Bolu 12.0 12.0 326
Turkey Slip
Duzce, Strike-
127 1605 7.14 1999 Duzce 6.6 0.0 276
Turkey Slip
Duzce, Strike-
128 1611 7.14 1999 Lamont 1058 0.2 0.2 425
Turkey Slip
Duzce, Strike-
129 1612 7.14 1999 Lamont 1059 4.2 4.2 425
Turkey Slip
Duzce, Strike-
130 1614 7.14 1999 Lamont 1061 11.5 11.5 481
Turkey Slip
Duzce, Strike-
131 1615 7.14 1999 Lamont 1062 9.2 9.2 338
Turkey Slip
Duzce, Strike-
132 1616 7.14 1999 Lamont 362 23.4 23.4 517
Turkey Slip
Duzce, Strike-
133 1617 7.14 1999 Lamont 375 3.9 3.9 425
Turkey Slip
Duzce, Strike-
134 1618 7.14 1999 Lamont 531 8.0 8.0 660
Turkey Slip
St Elias,
135 1628 7.54 1979 Reverse Icy Bay 26.5 26.5 275
Alaska
Strike-
136 1633 7.37 Manjil, Iran 1990 Abbar 12.6 12.6 724
Slip
Appendix A: MRCD 137 Ground Motion Set 321

Table A.2 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault ClstD -Boore Vs30
No. Rec. Year Station Name
(Mw) Name Type (km) Dist. (m/s)
No.
(km)
Strike-
137 1787 7.13 Hector Mine 1999 Hector 11.7 10.4 685
Slip
Appendix B

Description of Case Study Structures Used in Chapters 5-7

This appendix documents the generic moment-resisting frame (MRF), generic shear wall,
and reinforced concrete (RC) MRF structures and accompanying ground motion sets used
in Chapters 5-7. For more detailed information, the reader is referred to the
corresponding reports cited in the appropriate sections below.

B.1 Generic moment-resisting frame structures


The generic moment-resisting frame structures are taken from Zareian and Krawinkler
(2009, Chapter 4). The frames used were all 3-bay generic frames with story height equal
to 12 and bay span equal to 36. The models considered were two-dimensional centerline
models. The number of stories varied between 4 and 16 and for each number of stories,
three different fundamental periods were considered, equal to 0.1, 0.15 and 0.2 times the
number of stories. Viscous damping equal to 5% critical at the first and third mode of
vibration was assumed. For each period considered, three yield base shear coefficients ()
were considered. These coefficients were obtained by dividing a typical design response
spectrum corresponding to a 10/50 hazard level in the Los Angeles area and soil type D
by three values of R (1.5, 3 and 6). R corresponds to the ductility dependent reduction
factor. The column-to-beam bending strength was also varied and three different cases
were considered. Finally, the parameters of the component hysteretic behavior were also
varied, by selecting different values for each of the hinge plastic rotation, the post-
capping rotation capacity ration and capping strength ratio. Additional details are
provided by Zareian and Krawinkler (2009, Section 4.4).

The ground motion set considered was the LMSR-N, as developed by Medina and
Krawinkler (2003, Table 3.6). This set consists of 40 recordings where each record is
from a separate station (i.e., one random horizontal component is used from each station)
from events with magnitudes of Mw = 6.53 6.93 and distances of Rjb = 0.0 km

322
Appendix B: Description of Case Studies Used in Chapters 5-7 323

37.7 km. Additional details about the records in this set are provided by Medina and
Krawinkler (2003, Section 3.3).

B.2 Generic shear wall structures


The generic shear wall structures are also taken from Zareian and Krawinkler (2009,
Chapter 4). Only shear walls whose behavior is governed by flexural-type deformations
were considered. They were modeled as cantilevered members consisting of a series of
beam elements. Again, the number of stories varied between 4 and 16. However, the
corresponding fundamental periods were taken at 0.05, 0.075 and 0.1 times the number of
stories. Viscous damping was also equal to 5% critical at the first and third mode. The
same procedure that was followed to obtain the yield base shear coefficients in the case
of the moment resisting frame structures was also followed to obtain the bending strength
of the shear wall structures. The stiffness of the walls was not varied along the height,
consistent with practice in low- and mid-rise buildings. Finally, parameters of the
hysteretic behavior, including the plastic hinge rotation capacity, the post-capping
rotation capacity ratio, the capping strength ratio and the cyclic deterioration parameter
were varied based on existing literature. Additional details are provided by Zareian and
Krawinkler (2009, Section 4.5). The LMSR-N ground motion set was also used.

B.3 Reinforced concrete moment-resisting frame structures


The reinforced concrete moment-resisting frame structures were taken from Haselton and
Deierlein (2007, Table 3-3). All 30 frame structures were designed according to then-
current building codes and standards. The number of stories varied between 1 and 20,
with first mode periods between 0.42 s and 2.63 s. In all cases the story height was equal
to 12, while the bays were either 20 or 30 wide. Aside from frames with stepped
strength and stiffness along the height, as per common design practice, frames with
neither size nor reinforcement decrease along the height, as well as frames with a weak
story were considered. Boundary conditions also varied between completely fixed and
pinned, but only for the 1- and 2-story structures. All other structures were assumed to be
partially fixed by means of a grade beam. Additional details are provided by Haselton
and Deierlein (2007, Chapter 3).
Appendix B: Description of Case Studies Used in Chapters 5-7 324

A set of 39 recordings each with two horizontal components denoted Set One was
used. The set contains a maximum of six recordings per event. The records are from
events with magnitudes of Mw = 6.50 7.62 and distances of Rjb = 0.9 km 74.2 km.
Records have peak ground acceleration > 0.2 g and peak ground velocity > 15 cm/s.
Additional details are provided by Haselton and Deierlein (2007, Appendix 3B).
Appendix C

CS Ground Motion Sets Used in Chapters 6 and 7

This appendix documents the ground motions sets based on a conditional spectrum (CS)
that are used in Chapters 6 and 7. The CS describes both the expected spectral values and
the dispersion of those values at each period conditioned on the occurrence of a particular
event, known as the target event. As discussed by Lin et al. (2013a; 2013b), a CS
describes the full distribution of the spectrum at all periods (i.e., it captures both the
expected spectral values and the variability of those values) given the occurrence of an
event or events. The properties of an event (e.g., magnitude, distance, , fault type, etc.)
are used with a ground motion prediction equation (GMPE) and correlations between
values at different periods to construct the CS. Further information about the CS,
including how it can account for multiple events and multiple GMPEs, is provided by Lin
et al. (2013a; 2013b).

Two CS are used in Chapter 6 and are constructed using publicly available software,
which is described by Jayaram et al. (2011). This software also selects a user-specified
number of records from the PEER Next Generation Attenuation (NGA) database (Chiou
et al. 2008) that best fit the CS. The 2008 Boore and Atkinson (BA08) GMPE (Boore and
Atkinson 2008) is used for all calculations, and only one event is used for each CS. The
event properties are as follows: magnitude Mw = 6.7, Joyner-Boore distance Rjb = 10 km,
average soil velocity in the top 30 m of Vs30 = 285 m/s, and an unspecified fault type. One
CS is based on (1) = 1.7 while the other is based on (1) = 1.8, where T1 = 1.33 s.
Forty recordings, each with two horizontal components, are selected to match each CS.
The match is evaluated using the geometric mean spectral ordinates of each recording as
computed from its two horizontal components. Fourteen recordings are common to both
ground motion sets.

325
Appendix C: CS Ground Motion Sets Used in Chapters 6 and 7 326

C.1 CS, (1) = 1.7 ground motion set


As previously stated the CS, (1) = 1.7 ground motion set is based on the CS computed
using the BA08 GMPE with the following parameters: Mw = 6.7, Rjb = 10 km, Vs30 = 285
m/s, an unspecified fault type, and (1) = 1.7, where T1 = 1.33 s. The geometric mean
(i.e., the exponent of the mean of the logarithms) spectral values and the logarithmic
standard deviation of the CS are shown in Figure C.1(a) and Figure C.1(b), respectively.
Comparisons of how the selected ground motions compare to the target CS are also
included in these figures and demonstrate a close match between the two.

Figure C.1. Comparison of the target CS and the selected ground motions: (a) geometric
mean spectral values; (b) logarithmic standard deviation of spectral values.

The geometric mean spectrum of each selected ground motion is presented in Figure C.2
along with the median (geometric mean) spectrum and 2.5 and 97.5 percentiles of the
ground motion set.
Appendix C: CS Ground Motion Sets Used in Chapters 6 and 7 327

Figure C.2. Response spectra of ground motions in the CS, (T1)= 1.7 set.

Details about each of the forty recordings comprising the CS, (1) = 1.7 ground motion
set are presented in Table C.1, which includes the NGA identification numbers and
pertinent information about the events.

Table C.1. Records comprising the CS, (T1) = 1.7 ground motion set.
Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault Station ClstD -Boore Vs30
No. Rec. Year
(Mw) Name Type Name (km) Dist. (m/s)
No.
(km)
Imperial Strike- EC County
1 170 6.53 1979 7.3 7.3 192
Valley-06 Slip Center FF
2 Irpinia, Italy-
300 6.20 1980 Normal Calitri 8.8 8.8 600
02
Parkfield -
3 337 6.36 Coalinga-01 1983 Reverse Fault Zone 29.3 28.0 339
12
Parkfield -
4 348 6.36 Coalinga-01 1983 Reverse Gold Hill 36.2 35.0 339
1W
Parkfield -
5 359 6.36 Coalinga-01 1983 Reverse Vineyard 26.4 24.8 339
Cany 1E
6 Slack
369 6.36 Coalinga-01 1983 Reverse 27.5 26.0 685
Canyon
N. Palm Reverse- Morongo
7 527 6.06 1986 12.1 3.7 345
Springs Oblique Valley
Taiwan SMART1
8 571 7.30 1986 Reverse 0.0 0.0 275
SMART1(45) E01
9 Taiwan SMART1
583 7.30 1986 Reverse 0.0 0.0 275
SMART1(45) O10
Appendix C: CS Ground Motion Sets Used in Chapters 6 and 7 328

Table C.1 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault Station ClstD -Boore Vs30
No. Rec. Year
(Mw) Name Type Name (km) Dist. (m/s)
No.
(km)
Spitak, Reverse-
10 730 6.77 1988 Gukasian 0.0 0.0 275
Armenia Oblique
Bear Valley
11 Reverse- #12,
744 6.93 Loma Prieta 1989 51.0 50.7 331
Oblique Williams
Ranch
Reverse-
12 753 6.93 Loma Prieta 1989 Corralitos 3.9 0.2 462
Oblique
Reverse- Gilroy Array
13 766 6.93 Loma Prieta 1989 11.1 10.4 271
Oblique #2
Reverse- Gilroy Array
14 768 6.93 Loma Prieta 1989 14.3 13.8 222
Oblique #4
Olema -
Reverse-
15 785 6.93 Loma Prieta 1989 Point Reyes 117.1 117.0 339
Oblique
Station
16 Reverse- Treasure
808 6.93 Loma Prieta 1989 77.4 77.3 155
Oblique Island
Fountain
Strike-
17 856 7.28 Landers 1992 Valley - 146.9 146.9 270
Slip
Euclid
18 Strike-
864 7.28 Landers 1992 Joshua Tree 11.0 11.0 379
Slip
Beverly Hills
19 953 6.69 Northridge-01 1994 Reverse - 14145 17.2 9.4 356
Mulhol
LA -
20 Wadsworth
1009 6.69 Northridge-01 1994 Reverse 23.6 14.6 392
VA Hospital
North
Moorpark -
21 1039 6.69 Northridge-01 1994 Reverse 24.8 16.9 405
Fire Sta
Strike-
22 1116 6.90 Kobe, Japan 1995 Shin-Osaka 19.2 19.1 256
Slip
Strike-
23 1119 6.90 Kobe, Japan 1995 Takarazuka 0.3 0.0 312
Slip
Strike-
24 1120 6.90 Kobe, Japan 1995 Takatori 1.5 1.5 256
Slip
25 Kocaeli, Strike-
1166 7.51 1999 Iznik 30.7 30.7 275
Turkey Slip
Chi-Chi, Reverse-
26 1182 7.62 1999 CHY006 9.8 9.8 438
Taiwan Oblique
Chi-Chi, Reverse-
27 1187 7.62 1999 CHY015 38.1 38.1 229
Taiwan Oblique
28 Chi-Chi, Reverse-
1197 7.62 1999 CHY028 3.1 3.1 543
Taiwan Oblique
Chi-Chi, Reverse-
29 1227 7.62 1999 CHY074 10.8 0.7 553
Taiwan Oblique
Appendix C: CS Ground Motion Sets Used in Chapters 6 and 7 329

Table C.1 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault Station ClstD -Boore Vs30
No. Rec. Year
(Mw) Name Type Name (km) Dist. (m/s)
No.
(km)
Chi-Chi, Reverse-
30 1234 7.62 1999 CHY086 28.4 27.6 553
Taiwan Oblique
Chi-Chi, Reverse-
31 1286 7.62 1999 HWA037 46.2 41.7 273
Taiwan Oblique
32 Chi-Chi, Reverse-
1323 7.62 1999 ILA027 83.2 80.8 215
Taiwan Oblique
Chi-Chi, Reverse-
33 1330 7.62 1999 ILA039 86.1 83.8 227
Taiwan Oblique
Chi-Chi, Reverse-
34 1418 7.62 1999 TAP014 103.5 101.6 215
Taiwan Oblique
Chi-Chi, Reverse-
35 1421 7.62 1999 TAP021 101.4 99.5 215
Taiwan Oblique
Chi-Chi, Reverse-
36 1510 7.62 1999 TCU075 0.9 0.9 573
Taiwan Oblique
Chi-Chi, Reverse-
37 1536 7.62 1999 TCU110 11.6 11.6 213
Taiwan Oblique
Strike-
38 1633 7.37 Manjil, Iran 1990 Abbar 12.6 12.6 724
Slip
Chi-Chi,
39 2466 6.20 1999 Reverse CHY035 34.5 33.9 474
Taiwan-03
40 Chi-Chi,
3270 6.30 1999 Reverse CHY030 45.3 44.2 205
Taiwan-06

C.2 CS, (1) = 1.8 ground motion set


As previously stated the CS, (1) = 1.8 ground motion set is based on the CS computed
using the BA08 GMPE with the following parameters: Mw = 6.7, Rjb = 10 km, Vs30 = 285
m/s, an unspecified fault type, and (1) = 1.8, where T1 = 1.33 s. The geometric mean
(i.e., the exponent of the mean of the logarithms) spectral values and the logarithmic
standard deviation of the CS are shown in Figure C.3(a) and Figure C.3(b), respectively.
Comparisons of how the selected ground motions compare to the target CS are also
included in these figures and demonstrate a close match between the two.
Appendix C: CS Ground Motion Sets Used in Chapters 6 and 7 330

Figure C.3. Comparison of the target CS and the selected ground motions: (a) geometric
mean spectral values; (b) logarithmic standard deviation of spectral values.

The geometric mean spectrum of each selected ground motion is presented in Figure C.4
along with the median (geometric mean) spectrum and 2.5 and 97.5 percentiles of the
ground motion set.

Figure C.4. Response spectra of ground motions in the CS, (1) = 1.8 set.

Details about each of the forty recordings comprising the CS, (1) = 1.8 ground motion
set are presented in Table C.2, which includes the NGA identification numbers and
pertinent information about the events.
Appendix C: CS Ground Motion Sets Used in Chapters 6 and 7 331

Table C.2. Records comprising the CS, (T1) = 1.8 ground motion set.
Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault Station ClstD -Boore Vs30
No. Rec. Year
(Mw) Name Type Name (km) Dist. (m/s)
No.
(km)
Imperial Strike- EC County
1 170 6.53 1979 7.3 7.3 192
Valley-06 Slip Center FF
Irpinia, Italy-
2 300 6.20 1980 Normal Calitri 8.8 8.8 600
02
Parkfield -
3 337 6.36 Coalinga-01 1983 Reverse Fault Zone 29.3 28.0 339
12
N. Palm Reverse- Morongo
4 527 6.06 1986 12.1 3.7 345
Springs Oblique Valley
5 Taiwan SMART1
576 7.30 1986 Reverse 0.0 0.0 275
SMART1(45) M07
El Centro
Superstition Strike-
6 721 6.54 1987 Imp. Co. 18.2 18.2 192
Hills-02 Slip
Cent
APEEL 2 -
7 Reverse-
732 6.93 Loma Prieta 1989 Redwood 43.2 43.1 133
Oblique
City
Bear Valley
Reverse- #12,
8 744 6.93 Loma Prieta 1989 51.0 50.7 331
Oblique Williams
Ranch
9 Reverse-
779 6.93 Loma Prieta 1989 LGPC 3.9 0.0 478
Oblique
Larkspur
Reverse- Ferry
10 780 6.93 Loma Prieta 1989 94.6 94.6 170
Oblique Terminal
(FF)
Oakland -
Reverse-
11 783 6.93 Loma Prieta 1989 Outer Harbor 74.3 74.2 249
Oblique
Wharf
Strike-
12 864 7.28 Landers 1992 Joshua Tree 11.0 11.0 379
Slip
Beverly Hills
13 953 6.69 Northridge-01 1994 Reverse - 14145 17.2 9.4 356
Mulhol
Canoga Park
14 959 6.69 Northridge-01 1994 Reverse - Topanga 14.7 0.0 267
Can
Northridge -
15 1048 6.69 Northridge-01 1994 Reverse 17645 12.1 0.0 281
Saticoy St
Santa
16 1077 6.69 Northridge-01 1994 Reverse Monica City 26.5 17.3 336
Hall
17 Simi Valley -
1080 6.69 Northridge-01 1994 Reverse 13.4 0.0 557
Katherine Rd
Strike- Port Island (0
18 1114 6.90 Kobe, Japan 1995 3.3 3.3 198
Slip m)
Appendix C: CS Ground Motion Sets Used in Chapters 6 and 7 332

Table C.2 continued


Event Information Site Information
Joyner
NGA
Mag. Earthquake Fault Station ClstD -Boore Vs30
No. Rec. Year
(Mw) Name Type Name (km) Dist. (m/s)
No.
(km)
Strike-
19 1120 6.90 Kobe, Japan 1995 Takatori 1.5 1.5 256
Slip
20 1141 6.40 Dinar, Turkey 1995 Normal Dinar 3.4 0.0 220
Kocaeli, Strike-
21 1155 7.51 1999 Bursa Tofas 60.4 60.4 275
Turkey Slip
22 Chi-Chi, Reverse-
1178 7.62 1999 ALS 10.8 0.7 553
Taiwan Oblique
Chi-Chi, Reverse-
23 1182 7.62 1999 CHY006 9.8 9.8 438
Taiwan Oblique
Chi-Chi, Reverse-
24 1187 7.62 1999 CHY015 38.1 38.1 229
Taiwan Oblique
25 Chi-Chi, Reverse-
1197 7.62 1999 CHY028 3.1 3.1 543
Taiwan Oblique
Chi-Chi, Reverse-
26 1204 7.62 1999 CHY039 31.9 31.9 201
Taiwan Oblique
Chi-Chi, Reverse-
27 1209 7.62 1999 CHY047 24.1 24.1 273
Taiwan Oblique
28 Chi-Chi, Reverse-
1263 7.62 1999 HWA012 56.7 53.0 273
Taiwan Oblique
Chi-Chi, Reverse-
29 1286 7.62 1999 HWA037 46.2 41.7 273
Taiwan Oblique
Chi-Chi, Reverse-
30 1295 7.62 1999 HWA049 50.8 46.7 273
Taiwan Oblique
31 Chi-Chi, Reverse-
1323 7.62 1999 ILA027 83.2 80.8 215
Taiwan Oblique
Chi-Chi, Reverse-
32 1413 7.62 1999 TAP007 103.7 102.2 215
Taiwan Oblique
Chi-Chi, Reverse-
33 1421 7.62 1999 TAP021 101.4 99.5 215
Taiwan Oblique
34 Strike-
1602 7.14 Duzce, Turkey 1999 Bolu 12.0 12.0 326
Slip
St Elias,
35 1628 7.54 1979 Reverse Icy Bay 26.5 26.5 275
Alaska
Chi-Chi,
36 2461 6.20 1999 Reverse CHY028 24.4 23.4 543
Taiwan-03
Chi-Chi, Strike-
37 2734 6.20 1999 CHY074 6.2 6.0 553
Taiwan-04 Slip
Chi-Chi,
38 3080 6.20 1999 Reverse KAU020 109.1 106.0 373
Taiwan-05
Chi-Chi,
39 3467 6.30 1999 Reverse TCU065 26.1 24.1 306
Taiwan-06
Chi-Chi,
40 3510 6.30 1999 Reverse TCU139 39.0 37.7 304
Taiwan-06
List of References

Abrahamson, N., Atkinson, G., Boore, D., Bozorgnia, Y., Campbell, K., Chiou, B., Idriss,
I. M., Silva, W., and Youngs, R. (2008). "Comparisons of the NGA ground-motion
relations." Earthquake Spectra, 24(1), 45.

ACI (2005). Building code requirements for structural concrete (ACI 318-05) and
commentary (ACI 318R-05), American Concrete Institute (ACI), Farmington Hills,
MI.

Adam, C., and Jger, C. (2012a). "Seismic collapse capacity of basic inelastic structures
vulnerable to the P-delta effect." Earthquake Engineering & Structural Dynamics,
41(4), 775-793.

Adam, C., and Jger, C. (2012b). "Simplified collapse capacity assessment of earthquake
excited regular frame structures vulnerable to P-delta." Engineering Structures, 44,
159-173.

Agresti, A., and Coull, B. A. (1998). "Approximate is better than 'exact' for interval
estimation of binomial proportions." The American Statistician, 52(2), 119-126.

Agresti, A. (2002). Categorical data analysis, John Wiley & Sons, Inc., Hoboken, NJ.

AISC (2005a). "Prequalified connections for special and intermediate steel moment
frames for seismic applications, AISC 358-05." American Institute of Steel
Construction (AISC), Chicago, IL.

AISC (2005b). "Seismic provisions for structural steel buildings, including Supplement
No. 1." American Institute of Steel Construction (AISC), Chicago, IL.

Akkar, S., and Miranda, E. (2004). "Improved displacement modification factor to


estimate maximum deformations of short period structures." Proc., 13th World
Conference on Earthquake Engineering, Vancouver, BC, Canada, Paper No. 3424.

333
List of References 334

Akkar, S., and zen, . (2005). "Effect of peak ground velocity on deformation demands
for SDOF systems." Earthquake Engineering and Structural Dynamics, 34(13), 1551-
1571.

Akkar, S., Sandkkaya, M. A., and Bommer, J. J. (2013). "Empirical ground-motion


models for point- and extended-source crustal earthquake scenarios in Europe and the
Middle East." Bull Earthquake Eng, 1-29.

Alavi, B., and Krawinkler, H. (2001). "Effects of near-fault ground motions on frame
structures." Report No. 138, The John A. Blume Earthquake Engineering Center,
Stanford University, Stanford, CA, <http://blume.stanford.edu/tech_reports> (14
Aug. 2013).

Anderson, J. C., and Bertero, V. V. (1987). "Uncertainties in establishing design


earthquakes." Journal of Structural Engineering, 113(8).

ASCE (2005). Minimum design loads for buildings and other structures (ASCE/SEI 7-
05), American Society of Civil Engineers (ASCE), Reston, VA.

ASCE (2010). Minimum design loads for buildings and other structures (ASCE/SEI 7-
10), American Society of Civil Engineers (ASCE), Reston, VA.

Aslani, H., and Miranda, E. (2004). "Optimization of response simulation for loss
estimation using PEER's methodology." 13th World Conference on Earthquake
Engineering, Vancouver, Canada.

Aslani, H., and Miranda, E. (2005a). "Probabilistic earthquake loss estimation and loss
disaggregation in buildings." Report No. 157, The John A. Blume Earthquake
Engineering Center, Stanford University, Stanford, CA,
<http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Aslani, H., and Miranda, E. (2005b). "Probability-based seismic response analysis."


Engineering Structures, 27(8), 1151-1163.
List of References 335

ATC (2009a). "Quantification of building seismic performance factors." FEMA P695,


Federal Emergency Management Agency (FEMA), Washington, DC,
<http://www.fema.gov/media-library/assets/documents/16648?id=3736> (14 Aug.
2013).

ATC (2009b). "Effects of strength and stiffness degradation on seismic response." FEMA
P440A, Federal Emergency Management Agency (FEMA), Washington, DC,
<http://www.fema.gov/media-library/assets/documents/17037?id=3807> (14 Aug.
2013).

ATC (2012). "Seismic performance assessment of buildings." FEMA P-58, Federal


Emergency Management Agency (FEMA), Washington, DC,
<http://www.atcouncil.org/Projects/atc-58-project.html> (14 Aug. 2013).

Azarbakht, A., and Dolek, M. (2007). "Prediction of the median IDA curve by
employing a limited number of ground motion records." Earthquake Engineering and
Structural Dynamics, 36(15), 2401-2421.

Azarbakht, A., and Dolek, M. (2010). "Progressive incremental dynamic analysis for
first-mode dominated structures." Journal of Structural Engineering, 137(3), 445
455.

Baez, J. I., and Miranda, E. (2000). "Amplification factors to estimate inelastic


displacement demands for the design of structures in the near field." Proc., 12th
World Conference on Earthquake Engineering, Auckland, New Zealand.

Baker, J. W., and Cornell, C. A. (2005a). "Vector-valued ground motion intensity


measures for probabilistic seismic demand analysis." Report No. 150, The John. A
Blume Earthquake Engineering Center, Stanford University, Stanford, CA,
<http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Baker, J. W., and Cornell, C. A. (2005b). "A vector-valued ground motion intensity
measure consisting of spectral acceleration and epsilon." Earthquake Engineering &
Structural Dynamics, 34(10), 1193-1217.
List of References 336

Baker, J. W., and Cornell, C. A. (2006a). "Spectral shape, epsilon and record selection."
Earthquake Engineering and Structural Dynamics, 35(9), 1077-1095.

Baker, J. W., and Cornell, C. A. (2006b). "Which spectral acceleration are you using?"
Earthquake Spectra, 22(2), 293-312.

Baker, J. W. (2007a). "Measuring bias in structural response caused by ground motion


scaling." Proc., 8th Pacific Conference on Earthquake Engineering, Nanyang
Technological University, Singapore.

Baker, J. W. (2007b). "Quantitative classification of near-fault ground motions using


wavelet analysis." Bulletin of the Seismological Society of America, 97(5), 1486-
1501.

Baker, J. W. (2008). "An introduction to probabilistic seismic hazard analysis (PSHA),"


White Paper, Version 1.3,
<http://www.stanford.edu/~bakerjw/Publications/Baker_(2008)_Intro_to_PSHA_v1_
3.pdf> (14 Aug. 2013).

Baker, J. W., and Cornell, C. A. (2008). "Vector-valued intensity measures for pulse-like
near-fault ground motions." Engineering Structures, 30(4), 1048-1057.

Baker, J. W., and Jayaram, N. (2008). "Correlation of spectral acceleration values from
NGA ground motion models." Earthquake Spectra, 24(1), 299-317.

Baker, J. W. (2011). "Conditional mean spectrum: tool for ground-motion selection."


Journal of Structural Engineering, 137(3), 322-331.

Baker, J. W., Lin, T., Shahi, S. K., and Jayaram, N. (2011). "New ground motion
selection procedures and selected motions for the PEER transportation research
program." PEER Report 2011/03, Pacific Earthquake Engineering Research (PEER)
Center, Berkeley, CA,
<http://peer.berkeley.edu/publications/peer_reports/reports_2011/reports_2011.html>
(14 Aug. 2013).
List of References 337

Baker, J. W. (2013). "Efficient analytical fragility function fitting using dynamic


structural analysis." Earthquake Spectra, in review.

Bazzurro, P. (1998). "Probabilistic seismic demand analysis." PhD Dissertation,


Department of Civil and Environmental Engineering, Stanford University, Stanford,
CA, <http://blume.stanford.edu/rms_theses> (14 Aug. 2013).

Bazzurro, P., Cornell, C. A., Shome, N., and Carballo, J. (1998). "Three proposals for
characterizing MDOF nonlinear seismic response." Journal of Structural
Engineering, 124(11), 1281-1289.

Benjamin, J. R., and Cornell, C. A. (1970). Probability, statistics, and decision for civil
engineers, McGraw-Hill, New York, NY.

Bernal, D. (1992). "Instability of buildings subjected to earthquakes." Journal of


Structural Engineering, 118(8), 2239-2260.

Bertero, V. V., Herrera, R. A., and Mahin, S. A. (1976). "Establishment of design


earthquakes-evaluation of present methods." Proc., Int. Symp. on Earthquake
Structural Engineering, St. Louis, Missouri, 551-580.

Bertero, V. V., Mahin, S. A., and Herrera, R. A. (1978). "Aseismic design implications of
near-fault San Fernando earthquake records." Earthquake Engineering & Structural
Dynamics, 6(1), 31-42.

Bianchini, M. (2008). "Improved scalar intensity measures in performance-based


earthquake engineering." Nota Tecnica n. 215, Universit di Bologna, Bologna, Italy,
<http://amsacta.unibo.it/2455/1/Average_NT215.pdf> (14 Aug. 2013).

Bianchini, M., Diotallevi, P. P., and Baker, J. W. (2009). "Prediction of inelastic


structural response using an average of spectral accelerations." 10th International
Conference on Structural Safety and Reliability (ICOSSAR09), Osaka, Japan.

Blom, G. (1958). Statistical estimates and transformed beta variables, John Willey &
Sons, New York.
List of References 338

Bojrquez, E., and Iervolino, I. (2011). "Spectral shape proxies and nonlinear structural
response." Soil Dynamics and Earthquake Engineering, 31(7), 996-1008.

Boore, D. M., Watson-Lamprey, J., and Abrahamson, N. A. (2006). "Orientation-


independent measures of ground motion." Bulletin of the Seismological Society of
America, 96(4A), 1502-1511.

Boore, D. M., and Atkinson, G. M. (2008). "Ground-motion prediction equations for the
average horizontal component of PGA, PGV, and 5%-damped PSA at spectral
periods between 0.01s and 10.0s." Earthquake Spectra, 24(1), 99-138.

Bradley, B. A., and Dhakal, R. P. (2008). "Error estimation of closed-form solution for
annual rate of structural collapse." Earthquake Engineering and Structural Dynamics,
37(15), 1721-1737.

Bradley, B. A. (2010). "A generalized conditional intensity measure approach and


holistic ground-motion selection." Earthquake Engineering & Structural Dynamics,
39(12), 1321-1342.

Bradley, B. A., Dhakal, R. P., MacRae, G. A., and Cubrinovski, M. (2010a). "Prediction
of spatially distributed seismic demands in specific structures: ground motion and
structural response." Earthquake Engineering & Structural Dynamics, 39(5), 501-
520.

Bradley, B. A., Dhakal, R. P., MacRae, G. A., and Cubrinovski, M. (2010b). "Prediction
of spatially distributed seismic demands in specific structures: structural response to
loss estimation." Earthquake Engineering & Structural Dynamics, 39(6), 591-613.

Bradley, B. A. (2012). "The seismic demand hazard and importance of the conditioning
intensity measure." Earthquake Engineering & Structural Dynamics, 41(11), 1417-
1437.
List of References 339

BSSC (2003). "NEHRP recommended provisions for seismic regulations for new
buildings and other structures." FEMA 450, Federal Emergency Management Agency
(FEMA), Washington, DC, <http://www.fema.gov/media-
library/assets/documents/5543?id=2020> (14 Aug. 2013).

Burks, L. S., and Baker, J. W. (2012). "Occurrence of negative epsilon in seismic hazard
analysis deaggregation, and its impact on target spectra computation." Earthquake
Engineering & Structural Dynamics, 41(8), 1241-1256.

Campbell, K. W., and Bozorgnia, Y. (2008). "NGA ground motion model for the
geometric mean horizontal component of PGA, PGV, PGD and 5% damped linear
elastic response spectra for periods ranging from 0.01 to 10s." Earthquake Spectra,
24(1), 139.

Carballo, J. E., and Cornell, C. A. (2000). "Probabilistic seismic demand analysis:


spectrum matching and design." Report No. RMS-41, RMS Program, Stanford
University, Stanford, CA, <http://blume.stanford.edu/rms_theses> (14 Aug. 2013).

Challa, V. R. M., and Hall, J. F. (1994). "Earthquake collapse analysis of steel frames."
Earthquake Engineering and Structural Dynamics, 23(11), 1199-1218.

Chandramohan, R., Lin, T., Baker, J. W., and Deierlein, G. G. (2013). "Influence of
ground motion spectral shape and duration on seismic collapse risk." Proc., 10th
International Conference on Urban Earthquake Engineering, Tokyo, Japan.

Chiou, B., Darragh, R., Gregor, N., and Silva, W. (2008). "NGA project strong-motion
database." Earthquake Spectra, 24(1), 23-44.

Chopra, A. K., Bertero, V. V., and Mahin, S. A. (1973). "Response of the Olive View
Medical Center main building during the San Fernando earthquake." Proc., 5th World
Conference on Earthquake Engineering, Rome, Italy, 26-35.

Chopra, A. K. (2001). Dynamics of structures: theory and applications to earthquake


engineering, Prentice Hall, Upper Saddle River, New Jersey.
List of References 340

Chopra, A. K., and Goel, R. K. (2002). "A modal pushover analysis procedure for
estimating seismic demands for buildings." Earthquake Engineering and Structural
Dynamics, 31(3), 561-582.

Christovasilis, I. P., Filiatrault, A., Constantinou, M. C., and Wanitkorkul, A. (2009).


"Incremental dynamic analysis of woodframe buildings." Earthquake Engineering &
Structural Dynamics, 38(4), 477-496.

Clopper, C. J., and Pearson, E. S. (1934). "The use of confidence or fiducial limits
illustrated in the case of the binomial." Biometrika, 26, 404-413.

Cordova, P. P., Deierlein, G. G., Mehanny, S. S. F., and Cornell, C. A. (2000).


"Development of a two-parameter seismic intensity measure and probabilistic
assessment procedure." Proc., The Second U.S.-Japan Workshop on Performance-
Based Earthquake Engineering Methodology for Reinforced Concrete Building
Structures, Sapporo, Hokkaido, Japan, 187-206, 11-13 September, 2000.

Cuesta, I., and Aschheim, M. (2001). "Inelastic response spectra using conventional and
pulse R-factors." Journal of Structural Engineering, 127(9), 1013-1020.

Deierlein, G. G. (2004). "Overview of a comprehensive framework for earthquake


performance assessment." Proc., International Workshop on Performance-Based
Seismic Design - Concepts and Implementation, Bled, Slovenia, P. Fajfar and H.
Krawinkler, eds., Report No. PEER 2004/05, Berkeley, CA, 15-26,
<http://peer.berkeley.edu/publications/peer_reports/reports_2004/reports_2004.html>
(14 Aug. 2013).

Dolek, M. (2009). "Incremental dynamic analysis with consideration of modeling


uncertainties." Earthquake Engineering & Structural Dynamics, 38(6), 805-825.
List of References 341

Douglas, J. (2011). "Ground-motion prediction equations 1964-2010." PEER Report


2011/102, Pacific Earthquake Engineering Research Center (PEER), University of
California, Berkeley, CA,
<http://peer.berkeley.edu/publications/peer_reports/reports_2011/reports_2011.html>
(14 Aug. 2013).

Eads, L., Miranda, E., Krawinkler, H., and Lignos, D. G. (2012). "Deaggregation of
collapse risk." Proc., ASCE Structures Congress, Chicago, IL, March 29-31, 2012.

Eads, L., Miranda, E., Krawinkler, H., and Lignos, D. G. (2013). "An efficient method
for estimating the collapse risk of structures in seismic regions." Earthquake
Engineering & Structural Dynamics, 42(1), 25-41.

Efron, B. (1979). "Bootstrap methods: another look at the jackknife." The Annals of
Statistics, 7(1), 1-26.

Efron, B., and Tibshirani, R. J. (1993). An introduction to the bootstrap, Chapman &
Hall/CRC, New York.

Elkady, A., and Lignos, D. (2013). "Effect of composite action on the dynamic stability
of special steel moment resisting frames designed in seismic regions." Proc., ASCE
Structures Congress 2013, Pittsburgh, PA, 2151-2160.

Ellingwood, B. R., Celik, O. C., and Kinali, K. (2007). "Fragility assessment of building
structural systems in Mid-America." Earthquake Engineering & Structural
Dynamics, 36(13), 1935-1952.

Fajfar, P., and Dolek, M. (2012). "A practice-oriented estimation of the failure
probability of building structures." Earthquake Engineering & Structural Dynamics,
41(3), 531-547.

Field, E. H., Jordan, T. H., and Cornell, C. A. (2003). "A developing community-
modeling environment for seismic hazard analysis." Seismological Research Letters,
74(4), 406-419.
List of References 342

Field, E. H. (2013). "OpenSHA." OpenSHA.org and the University of Southern


California, <http://www.opensha.org> (14 Aug. 2013).

Ger, J.-F., Cheng, F. Y., and Lu, L.-W. (1993). "Collapse behavior of Pino Suarez
building during 1985 Mexico City earthquake." Journal of Structural Engineering
(ASCE), 119(3), 852-870.

Ghafory-Ashtiany, M., Mousavi, M., and Azarbakht, A. (2011). "Strong ground motion
record selection for the reliable prediction of the mean seismic collapse capacity of a
structure group." Earthquake Engineering and Structural Dynamics, 40(6), 691-708.

Gregor, N. J., Addo, K. O., Abrahamson, N. A., and Youngs, R. R. (2013). "Comparison
of BC Hydro Subduction GMPE to data from recent large megathrust earthquakes."
Proc., 15th World Conference on Earthquake Engineering, Lisbon, Portugal.

Guamn, J. W. (2010). "Empirical ground motion relationships for maximum incremental


velocity." Master's Thesis, Department of Civil Engineering and Geological Sciences,
University of Notre Dame, Notre Dame, IN, <http://etd.nd.edu/ETD-
db/theses/available/etd-04162010-151559> (14 Aug. 2013).

Gupta, A., and Krawinkler, H. (1999). "Seismic demands for performance evaluation of
steel moment resisting frame structures." Report No. 132, The John A. Blume
Earthquake Engineering Center, Stanford University, Stanford, CA,
<http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Han, S. W., and Chopra, A. K. (2006). "Approximate incremental dynamic analysis using
the modal pushover analysis procedure." Earthquake Engineering & Structural
Dynamics, 35(15), 1853-1873.

Han, S. W., Moon, K.-H., and Chopra, A. K. (2010). "Application of MPA to estimate
probability of collapse of structures." Earthquake Engineering and Structural
Dynamics, 39(11), 1259-1278.
List of References 343

Haselton, C. B., and Baker, J. W. (2006). "Ground motion intensity measures for collapse
capacity prediction: choice of optimal spectral period and effect of spectral shape."
Proc., 8th National Conference on Earthquake Engineering, 18-22.

Haselton, C. B., and Deierlein, G. G. (2007). "Assessing seismic collapse safety of


modern reinforced concrete moment frame buildings." Report No. 156, The John A.
Blume Earthquake Engineering Center, Stanford University, Stanford, CA,
<http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Haselton, C. B., and Deierlein, G. G. (2008). "Assessing seismic collapse safety of


modern reinforced concrete moment-frame buildings." PEER Report 2007/08, Pacific
Earthquake Engineering Research Center (PEER), University of California, Berkeley,
CA,
<http://peer.berkeley.edu/publications/peer_reports/reports_2007/reports_2007.html>
(14 Aug. 2013).

Haselton, C. B., Goulet, C. A., Mitrani-Reiser, J., Beck, J. L., Deierlein, G. G., Porter, K.
A., Stewart, J. P., and Taciroglu, E. (2008). "An assessment to benchmark the seismic
performance of a code-conforming reinforced concrete moment-frame building."
PEER Report 2007/12, Pacific Earthquake Engineering Research Center (PEER),
University of California, Berkeley, CA,
<http://peer.berkeley.edu/publications/peer_reports/reports_2007/reports_2007.html>
(14 Aug. 2013).

Haselton, C. B., Baker, J. W., Liel, A. B., and Deierlein, G. G. (2011). "Accounting for
ground-motion spectral shape characteristics in structural collapse assessment through
an adjustment for epsilon." Journal of Structural Engineering, 137, 332-344.

Hines, E. M., Appel, M. E., and Cheever, P. J. (2009). "Collapse performance of low-
ductility chevron braced steel frames in moderate seismic regions." AISC Engineering
Journal, 46(3), 149-180.

Housner, G. W. (1952). "Spectrum intensities of strong-motion earthquakes." Proc.,


Symposium on Earthquake and Blast Effects on Structures, Los Angeles, CA, 21-36.
List of References 344

Hsiao, P.-C., Lehman, D. E., and Roeder, C. W. (2013). "Evaluation of the response
modification coefficient and collapse potential of special concentrically braced
frames." Earthquake Engineering & Structural Dynamics, 42(10), 1547-1564.

Huang, Y.-N. (2008). "Performance assessment of conventional and base-isolated nuclear


power plants for earthquake and blast loadings." PhD Dissertation, Department of
Civil, Structural and Environmental Engineering, State University of New York at
Buffalo, Buffalo, NY.

Ibarra, L., and Krawinkler, H. (2011). "Variance of collapse capacity of SDOF systems
under earthquake excitations." Earthquake Engineering & Structural Dynamics,
40(12), 1299-1314.

Ibarra, L. F., Medina, R. A., and Krawinkler, H. (2002). "Collapse assessment of


deteriorating SDOF systems." Proc., 12th European Conference on Earthquake
Engineering, London, UK, Paper 665, Elsevier Science Ltd., Sept. 9-13, 2002.

Ibarra, L. F. (2003). "Global collapse of frame structures under seismic excitations." PhD
Dissertation, Department of Civil and Environmental Engineering, Stanford
University, Stanford, CA.

Ibarra, L. F., and Krawinkler, H. (2005). "Global collapse of frame structures under
seismic excitations." Report No. 152, The John A. Blume Earthquake Engineering
Center, Stanford University, Stanford, CA, <http://blume.stanford.edu/tech_reports>
(14 Aug. 2013).

Ibarra, L. F., Medina, R. A., and Krawinkler, H. (2005). "Hysteretic models that
incorporate strength and stiffness deterioration." Earthquake Engineering and
Structural Dynamics, 34(12), 1489-1511.

ICC (2003). International building code, International Code Council (ICC), Falls Church,
VA.

ICC (2005). International building code, International Code Council (ICC), Falls Church,
VA.
List of References 345

Jacobsen, L. S., and Ayre, R. S. (1952). "A comparative study of pulse and step-type
loads on a simple vibratory system." Technical Report No. 16, School of Engineering,
Stanford University, Stanford, CA.

Jalayer, F. (2003). "Direct probabilistic seismic analysis: implementing non-linear


dynamic assessments." PhD Dissertation, Department of Civil and Environmental
Engineering, Stanford University, Stanford, CA,
<http://blume.stanford.edu/rms_theses> (14 Aug. 2013).

Jalayer, F., Beck, J. L., Porter, K. A., and Hall, J. F. (2004). "Effects of ground motion
uncertainty on predicting the response of an existing RC frame structure." 13th World
Conference on Earthquake Engineering, Vancouver, Canada.

Jayaram, N., and Baker, J. W. (2008). "Statistical tests of the joint distribution of spectral
acceleration values." Bulletin of the Seismological Society of America, 98(5), 2231-
2243.

Jayaram, N., Lin, T., and Baker, J. W. (2011). "A computationally efficient ground-
motion selection algorithm for matching a target response spectrum mean and
variance." Earthquake Spectra, 27(3), 797-815.

Jennings, P. C., and Husid, R. (1968). "Collapse of yielding structures during


earthquakes." Journal of Engineering Mechanics Division (ASCE), 94(5), 1045-1065.

Kadas, K., Yakut, A., and Kazaz, I. (2011). "Spectral ground motion intensity based on
capacity and period elongation." Journal of Structural Engineering, 137(3), 9.

Kanvinde, A. M. (2003). "Methods to evaluate the dynamic stability of structures-shake


table tests and nonlinear dynamic analyses." Proc., 2003 EERI Annual Meeting, EERI
Paper Competition Winner, Portland, OR.
List of References 346

Karamanci, E. (2013). "Collapse assessment and performance-based evaluation


techniques for concentrically braced frames designed in seismic regions." M.Eng
Thesis, Department of Civil Engineering and Applied Mechanics, McGill University,
Montreal, Canada, <http://digitool.library.mcgill.ca/thesisfile117045.pdf> (14 Aug.
2013).

Kennedy, R. P., Short, S. A., Merz, K. L., Tokarz, F. J., Idriss, I. M., Power, M. S., and
Sadigh, K. (1984). "Engineering characterization of ground motion. Task I: Effects of
characteristics of free-field motion on structural response." Report No. NUREG/CR-
3805, U.S. Nuclear Regulatory Commission, Washington, DC,
<http://www.osti.gov/scitech/biblio/6848574> (14 Aug. 2013).

Krawinkler, H., and Miranda, E. (2004). "Performance-based earthquake engineering."


Earthquake engineering: from engineering seismology to performance-based
engineering, Y. Bozorgnia and V. V. Bertero, eds., CRC Press, Boca Raton, FL.

Krishnan, S., and Muto, M. (2012). "Mechanism of collapse of tall steel moment-frame
buildings under earthquake excitation." Journal of Structural Engineering, 138(11),
1361-1387.

Kurama, Y. C., and Farrow, K. T. (2003). "Ground motion scaling methods for different
site conditions and structure characteristics." Earthquake Engineering and Structural
Dynamics, 32(15), 2425-2450.

Lavan, O., Sivaselvan, M. V., Reinhorn, A. M., and Dargush, G. F. (2009). "Progressive
collapse analysis through strength degradation and fracture in the Mixed Lagrangian
Formulation." Earthquake Engineering & Structural Dynamics, 38(13), 1483-1504.

Lee, K., and Foutch, D. A. (2002). "Performance evaluation of new steel frame buildings
for seismic loads." Earthquake Engineering & Structural Dynamics, 31(3), 653-670.

Li, G., and Fahnestock, L. (2013). "Seismic response of single-degree-of-freedom


systems representing low-ductility steel concentrically braced frames with reserve
capacity." Journal of Structural Engineering, 139(2), 199-211.
List of References 347

Li, Y., and Ellingwood, B. R. (2007). "Reliability of woodframe residential construction


subjected to earthquakes." Structural Safety, 29(4), 294-307.

Liel, A. B., and Deierlein, G. G. (2008). "Assessing the collapse risk of Californias
existing reinforced concrete frame structures: metrics for seismic safety decisions."
Report No. 166, The John A. Blume Earthquake Engineering Center, Stanford
University, Stanford, CA, <http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Liel, A. B., Haselton, C. B., Deierlein, G. G., and Baker, J. W. (2009). "Incorporating
modeling uncertainties in the assessment of seismic collapse risk of buildings."
Structural Safety, 31(2), 197-211.

Liel, A. B., Haselton, C. B., and Deierlein, G. G. (2011). "Seismic collapse safety of
reinforced concrete buildings. II: Comparative assessment of nonductile and ductile
moment frames." Journal of Structural Engineering, 137(4), 492-502.

Lignos, D. G., and Krawinkler, H. (2007). "A database in support of modeling of


component deterioration for collapse prediction of steel frame structures." ASCE
Structures Congress, Long Beach, CA.

Lignos, D. G. (2009). IIIDAP: Interactive interface for incremental dynamic analysis,


Version 1.1.5, Department of Civil and Environmental Engineering, Stanford
University, <http://dimitrios-lignos.research.mcgill.ca/IIIDAP.html> (24 Feb. 2011).

Lignos, D. G., Eads, L., and Krawinkler, H. (2011a). "Effect of composite action on
collapse capacity of steel moment frames under earthquake loading." Proc.,
Eurosteel: 6th European Conference on Steel and Composite Structures, Budapest,
Hungary, Aug. 31 - Sept. 2, 2011.

Lignos, D. G., and Krawinkler, H. (2011). "Deterioration modeling of steel components


in support of collapse prediction of steel moment frames under earthquake loading."
Journal of Structural Engineering, 137(11), 1291-1302.
List of References 348

Lignos, D. G., Krawinkler, H., and Whittaker, A. S. (2011b). "Prediction and validation
of sidesway collapse of two scale models of a 4-story steel moment frame."
Earthquake Engineering and Structural Dynamics, 40(7), 807-825.

Lignos, D. G., and Krawinkler, H. (2012a). "Sidesway collapse of deteriorating structural


systems under seismic excitations." Report No. 177, The John A. Blume Earthquake
Engineering Center, Stanford University, Stanford, CA,
<http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Lignos, D. G., and Krawinkler, H. (2012b). "Development and utilization of structural


component databases for performance-based earthquake engineering." Journal of
Structural Engineering, 139(8), 1382-1394.

Lignos, D. G., Hikino, T., Matsuoka, Y., and Nakashima, M. (2013). "Collapse
assessment of steel moment frames based on E-Defense full-scale shake table
collapse tests." Journal of Structural Engineering, 139(1), 120-132.

Lignos, D. G., and Karamanci, E. (2013). "Predictive equations for modeling cyclic
buckling and fracture of steel braces." 10th International Conference on Urban
Earthquake Engineering, Tokyo, Japan.

Lin, T. (2012). "Advancement of hazard-consistent ground motion selection


methodology." PhD Dissertation, Department of Civil and Environmental
Engineering, Stanford University, Stanford, CA.

Lin, T., Harmsen, S. C., Baker, J. W., and Luco, N. (2013a). "Conditional spectrum
computation incorporating multiple causal earthquakes and ground motion prediction
models." Bulletin of the Seismological Society of America, 103(2A), 1103-1116.

Lin, T., Haselton, C. B., and Baker, J. W. (2013b). "Conditional spectrum-based ground
motion selection. Part I: Hazard consistency for risk-based assessments." Earthquake
Engineering & Structural Dynamics, (in press).
List of References 349

Luco, N., and Bazzurro, P. (2007). "Does amplitude scaling of ground motion records
result in biased nonlinear structural drift responses?" Earthquake Engineering &
Structural Dynamics, 36(13), 1813-1835.

Luco, N., and Cornell, C. A. (2007). "Structure-specific scalar intensity measures for
near-source and ordinary earthquake ground motions." Earthquake Spectra, 23(2),
357-392.

Luco, N., Ellingwood, B., Hamburger, R., Hooper, J., Kimball, J., and Kircher, C. (2007).
"Risk-targeted versus current seismic design maps for the conterminous United
States." Proc., SEAOC 2007 Convention, Squaw Creek, California, 163-175, Sept.
26-29, 2007.

Makris, N., and Black, C. J. (2004). "Evaluation of peak ground velocity as a good
intensity measure for near-source ground motions." Journal of Engineering
Mechanics, 130(9), 1032-1044.

Mavroeidis, G., and Papageorgiou, A. (2003). "A mathematical representation of near-


fault ground motions." Bulletin of the Seismological Society of America, 93(3), 1099.

McKenna, F. T. (1997). "Object-oriented finite element programming: frameworks for


analysis, algorithms and parallel computing." PhD Dissertation, Department of Civil
and Environmental Engineering, University of California, Berkeley, CA,
<http://opensees.berkeley.edu/OpenSees/doc/fmkdiss.pdf> (14 Aug. 2013).

McKenna, F. T. (2009). Open System for Earthquake Engineering Simulation


(OpenSees), Version 2.1.1, Pacific Earthquake Engineering Research Center
(PEER), <http://opensees.berkeley.edu> (6 June 2010).

Medina, R. A., and Krawinkler, H. (2003). "Seismic demands for nondeteriorating frame
structures and their dependence on ground motions." Report No. 144, The John A.
Blume Earthquake Engineering Center, Stanford University, Stanford, CA,
<http://blume.stanford.edu/tech_reports> (14 Aug. 2013).
List of References 350

Mehanny, S. S. F., and Deierlein, G. G. (2000). "Modeling and assessment of seismic


performance of composite frames with reinforced concrete columns and steel beams."
Report No. 135, The John A. Blume Earthquake Engineering Center, Stanford
University, Stanford, CA, <http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Menun, C., and Fu, Q. (2002). "An analytical model for near-fault ground motions and
the response of SDOF systems." Proc., 7th US National Conference on Earthquake
Engineering, Boston, Massachusetts.

Miranda, E. (2001). "Estimation of inelastic deformation demands of SDOF systems."


Journal of Structural Engineering, 127(9), 1005-1012.

Miranda, E., and Akkar, S. D. (2003). "Dynamic instability of simple structural systems."
Journal of Structural Engineering, 129(12), 1722-1726.

Mousavi, M., Ghafory-Ashtiany, M., and Azarbakht, A. (2011). "A new indicator of
elastic spectral shape for the reliable selection of ground motion records." Earthquake
Engineering & Structural Dynamics, 40(12), 1403-1416.

Newmark, N. M. (1959). "A method of computation for structural dynamics." Proc.


ASCE, 85(3), 67-94.

Newmark, N. M. (1973). "A study of vertical and horizontal earthquake spectra." Report
No. WASH-1255, US Atomic Energy Commission, Washington, DC.

NIST (2010). "Evaluation of the FEMA P-695 methodology for quantification of


building seismic performance factors." NIST GCR 10-917-8, prepared by the NEHRP
Consultants Joint Venture for the National Institute of Standards and Technology
(NIST), Gaithersburg, MD.

NIST (2011). "Selecting and scaling earthquake ground motions for performing response-
history analyses." NIST GCR 11-917-15, prepared by the NEHRP Consultants Joint
Venture for the National Institute of Standards and Technology (NIST), Gaithersburg,
Maryland.
List of References 351

Ordaz, M., and Prez-Rocha, L. E. (1998). "Estimation of strength-reduction factors for


elastoplastic systems: a new approach." Earthquake Engineering & Structural
Dynamics, 27(9), 889-901.

Ordaz, M., Martinelli, F., DAmico, V., and Meletti, C. (2013). "CRISIS2008: a flexible
tool to perform probabilistic seismic hazard assessment." Seismological Research
Letters, 84(3), 495-504.

Osteraas, J. D., and Krawinkler, H. (1990). "Strength and ductility considerations in


seismic design." Report No. 90, The John A. Blume Earthquake Engineering Center,
Stanford University, Stanford, CA, <http://blume.stanford.edu/tech_reports> (14
Aug. 2013).

PEER (2013). "NGA-West2 final products." Pacific Earthquake Engineering Research


(PEER) Center, University of California, Berkeley, CA,
<http://peer.berkeley.edu/ngawest2/final-products> (14 Aug. 2013).

Petersen, M. D., Frankel, A. D., Harmsen, S. C., Mueller, C. S., Haller, K. M., Wheeler,
R. L., Wesson, R. L., Zeng, Y., Boyd, O. S., Perkins, D. M., Luco, N., Field, E. H.,
Wills, C. J., and Rukstales, K. S. (2008). "Documentation for the 2008 update of the
United States National Seismic Hazard Maps." Open-File Report 20081128, United
States Geological Survey (USGS), Reston, VA, <http://pubs.usgs.gov/of/2008/1128>
(14 Aug. 2013).

Power, M., Chiou, B., Abrahamson, N., Bozorgnia, Y., Shantz, T., and Roblee, C. (2008).
"An overview of the NGA project." Earthquake Spectra, 24(1), 3-21.

Power, M. S., Youngs, R. R., and Chin, C.-C. (2007). "Design ground motion library."
Report of PEER-LL Program Task 1F01, Pacific Earthquake Engineering Research
(PEER) Center, University of California, Berkeley, CA,
<http://peer.berkeley.edu/lifelines/lifelines_pre_2006/final_reports/1F01-FR.pdf> (14
Aug. 2013).
List of References 352

Riddell, R. (2007). "On ground motion intensity indices." Earthquake Spectra, 23(1),
147-173.

Rubinstein, R. Y., and Kroese, D. P. (2008). Simulation and the Monte Carlo method,
2nd edition, John Wiley & Sons, Inc., Hoboken, NJ.

Ruiz-Garca, J., and Miranda, E. (2005). "Perfomance-based assessment of existing


structures accounting for residual displacements." Report No. 153, The John A.
Blume Earthquake Engineering Center, Stanford University, Stanford, CA,
<http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Scott, M., and Fenves, G. (2010). "Krylov subspace accelerated newton algorithm:
application to dynamic progressive collapse simulation of frames." Journal of
Structural Engineering, 136(5), 473-480.

Seed, H. B., Murarka, R., Lysmer, J., and Idriss, I. M. (1976). "Relationships of
maximum acceleration, maximum velocity, distance from source, and local site
conditions for moderately strong earthquakes." Bulletin of the Seismological Society
of America, 66(4), 1323-1342.

Sewell, R. T., Toro, G. R., and McGuire, R. K. (1996). "Impact of ground motion
characterization on conservatism and variability in seismic risk estimates." Report
No. NUREG/CR--6467, U.S. Nuclear Regulatory Commission, Washington, DC,
<http://www.osti.gov/scitech/biblio/285473> (14 Aug. 2013).

Shafei, B., Zareian, F., and Lignos, D. G. (2011). "A simplified method for collapse
capacity assessment of moment-resisting frame and shear wall structural systems."
Engineering Structures, 33(4), 1107-1116.

Shakib, H., and Pirizadeh, M. (2013). "Probabilistic seismic performance assessment of


setback buildings under bidirectional excitation." Journal of Structural Engineering,
posted ahead of print February 28, 2013.

Shome, N., Cornell, C. A., Bazzurro, P., and Carballo, J. E. (1998). "Earthquakes,
records, and nonlinear responses." Earthquake Spectra, 14(3), 469-500.
List of References 353

Shome, N., and Cornell, C. A. (1999). "Probabilistic seismic demand analysis of


nonlinear structures." Report No. RMS-35, RMS Program, Stanford University,
Stanford, CA, <http://blume.stanford.edu/rms-reports> (14 Aug. 2013).

Sivaselvan, M., and Reinhorn, A. (2002). "Collapse analysis: large inelastic deformations
analysis of planar frames." Journal of Structural Engineering, 128(12), 1575-1583.

Sivaselvan, M., and Reinhorn, A. (2006). "Lagrangian approach to structural collapse


simulation." Journal of Engineering Mechanics, 132(8), 795-805.

Sivaselvan, M. V., Lavan, O., Dargush, G. F., Kurino, H., Hyodo, Y., Fukuda, R., Sato,
K., Apostolakis, G., and Reinhorn, A. M. (2009). "Numerical collapse simulation of
large scale structural systems using an optimization based algorithm." Earthquake
Engineering and Structural Dynamics, 38(5), 655-677.

Song, J., and Ellingwood, B. (1999a). "Seismic reliability of special moment steel frames
with welded connections: I." Journal of Structural Engineering, 125(4), 357-371.

Song, J., and Ellingwood, B. (1999b). "Seismic reliability of special moment steel frames
with welded connections: II." Journal of Structural Engineering, 125(4), 372-384.

Stoakes, C. D. (2012). "Beam-column connection flexural behavior and seismic collapse


performance of concentrically braced frames." PhD Dissertation, Department of Civil
and Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana,
IL.

Sucuoglu, H., Erberik, M., and Yucemen, M. (1999). "Influence of peak ground velocity
on seismic failure probability." Proc., Proc. 4th Int. Conf. of the European
Association for Structural Dynamics (EURODYN99), Prague, Czech Republic.

Sun, C.-K., Berg, G. V., and Hanson, R. D. (1973). "Gravity effect on single-degree
inelastic system." Journal of the Engineering Mechanics Division, 99(1), 183-200.
List of References 354

Taghavi, S., and Miranda, E. (2006). "Probabilistic seismic assessment of floor


acceleration demands in multi-story buildings." Report No. 162, The John A. Blume
Earthquake Engineering Center, Stanford University, Stanford, CA,
<http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Takizawa, H., and Jennings, P. C. (1980). "Collapse of a model for ductile reinforced
concrete frames under extreme earthquake motions." Earthquake Engineering and
Structural Dynamics, 8(2), 117-144.

Tothong, P., and Cornell, C. A. (2006). "An empirical ground-motion attenuation relation
for inelastic spectral displacement." Bulletin of the Seismological Society of America,
96(6), 2146-2164.

Tothong, P., and Luco, N. (2007). "Probabilistic seismic demand analysis using advanced
ground motion intensity measures." Earthquake Engineering & Structural Dynamics,
36(13), 1837-1860.

Trifunac, M. D. (2009). "75th anniversary of strong motion observationa historical


review." Soil Dynamics and Earthquake Engineering, 29(4), 591-606.

Tsantaki, S., Jger, C., and Adam, C. (2012). "Improved seismic collapse prediction of
inelastic simple systems vulnerable to the P-delta effect based on average spectral
acceleration." 15th World Conference on Earthquake Engineering, Lisbon, Portugal.

Tsantaki, S., and Adam, C. (2013). "Collapse capacity spectra based on an improved
intensity measure." COMPDYN 2013: 4th International Conference on
Computational Methods in Structural Dynamics and Earthquake Engineering, Kos
Island, Greece.

Tso, W. K., Zhu, T. J., and Heidebrecht, A. C. (1992). "Engineering implication of


ground motion A/V ratio." Soil Dynamics and Earthquake Engineering, 11(3), 133-
144.

UBC (1994). Uniform Building Code, International Conference of Building Officials,


Whittier, CA.
List of References 355

USGS (2012a). "Hazard Curve Application." United States Geological Survery (USGS),
<http://geohazards.usgs.gov/hazardtool> (14 March 2013).

USGS (2012b). "2008 Interactive Deaggregations." United States Geological Survey


(USGS), <http://geohazards.usgs.gov/deaggint/2008> (14 March 2013).

Vamvatsikos, D. (2002). "Seismic performance, capacity and reliability of structures as


seen through incremental dynamic analysis." PhD Dissertation, Department of Civil
and Environmental Engineering, Stanford University, Stanford, CA,
<http://blume.stanford.edu/rms_theses> (14 Aug. 2013).

Vamvatsikos, D., and Cornell, C. A. (2004). "Applied incremental dynamic analysis."


Earthquake Spectra, 20(2), 523-553.

Vamvatsikos, D., and Cornell, C. A. (2005a). "Seismic performance, capacity and


reliability of structures as seen through Incremental Dynamic Analysis." Report No.
151, The John A. Blume Earthquake Engineering Center, Stanford University,
Stanford, CA, <http://blume.stanford.edu/tech_reports> (14 Aug. 2013).

Vamvatsikos, D., and Cornell, C. A. (2005b). "Developing efficient scalar and vector
intensity measures for IDA capacity estimation by incorporating elastic spectral shape
information." Earthquake Engineering & Structural Dynamics, 34(13), 1573-1600.

Vamvatsikos, D., and Cornell, C. A. (2006). "Direct estimation of the seismic demand
and capacity of oscillators with multi-linear static pushovers through IDA."
Earthquake Engineering & Structural Dynamics, 35(9), 1097-1117.

Vamvatsikos, D., and Fragiadakis, M. (2010). "Incremental dynamic analysis for


estimating seismic performance sensitivity and uncertainty." Earthquake Engineering
& Structural Dynamics, 39(2), 141-163.

Vamvatsikos, D. (2012). "Accurate application and higher-order solutions of the


SAC/FEMA probabilistic format for performance assessment." Proc., 15th World
Conference on Earthquake Engineering, Lisbon, Portugal.
List of References 356

Vian, D., and Bruneau, M. (2003). "Tests to structural collapse of single degree of
freedom frames subjected to earthquake excitations." Journal of Structural
Engineering, 129(12), 1676-1685.

Villaverde, R. (2007). "Methods to assess the seismic collapse capacity of building


structures: state of the art." Journal of Structural Engineering (ASCE), 133(1), 57-66.

Wang, T., Mosqueda, G., Jacobsen, A., and Cortes-Delgado, M. (2012). "Performance
evaluation of a distributed hybrid test framework to reproduce the collapse behavior
of a structure." Earthquake Engineering & Structural Dynamics, 41(2), 295-313.

Wen, Y. K. (1995). "Building reliability and code calibration." Earthquake Spectra,


11(2), 269-296.

Wilson, E. B. (1927). "Probable inference, the law of succession, and statistical


inference." Journal of the American Statistical Association, 22(158), 209-212.

Zareian, F., and Krawinkler, H. (2007a). "Assessment of probability of collapse and


design for collapse safety." Earthquake Engineering and Structural Dynamics,
36(13), 1901-1914.

Zareian, F., and Krawinkler, H. (2007b). "Sensitivity of collapse potential of buildings to


variations in structural systems and structural parameters." Structures Congress 2007,
Long Beach, CA.

Zareian, F., and Krawinkler, H. (2009). "Simplified performance-based earthquake


engineering." Report No. 169, The John A. Blume Earthquake Engineering Center,
Stanford University, Stanford, CA, <http://blume.stanford.edu/tech_reports> (14
Aug. 2013).

Zareian, F., and Medina, R. A. (2010). "A practical method for proper modeling of
structural damping in inelastic plane structural systems." Computers & Structures,
88(1-2), 45-53.
List of References 357

Zhu, T. J., Heidebrecht, A. C., and Tso, W. K. (1988). "Effect of peak ground
acceleration to velocity ratio on ductility demand of inelastic systems." Earthquake
Engineering & Structural Dynamics, 16(1), 63-79.

You might also like