You are on page 1of 278

AN ABSTRACT OF THE THESIS OF

Andre Belejo for the degree of Doctor of Philosophy in Civil Engineering presented on
September 14, 2017.

Title: Evaluation of ground motion duration effects on the damage prediction of


building and bridge structural and soil-structural systems

Abstract approved:
______________________________________________________
Andre R. Barbosa

The role of ground motion duration on the seismic performance of civil engineering
structures remains unclear. Thus, the role of earthquake strong ground motion duration
is assessed in a three manuscript thesis, which includes: (1) assessment of the effects
of ground motion duration on the seismic performance of a model of a three-
dimensional (3-D) plan-asymmetric reinforced concrete building tested in Europe; (2)
determination of the influence of ground motion duration and the impact of the retrofit
solution of columns using titanium alloy bars employed on vintage bridges built in the
1950s – 1970s in the western United States; and (3) estimation of the effect of soil-
structure interaction (SSI) on the assessment of ground motion duration of two steel
moment resisting frame buildings.
In this thesis, nonlinear finite element models of the building and bridge structures are
developed in OpenSees and validated using experimental data when these were
available. Then, the models are subjected to long- and short-duration earthquake
ground motions sets, in which each short- and long-duration ground motion pair have
similar spectral shapes. In the process, a new ground motion selection procedure is
proposed to isolate the effect of duration on the structural response of 3-D structures
subjected by bi-directional ground motions.
Damage index-based fragility curves are proposed in this work. These are based on
damage assessment of the structures analyzing global and local responses estimated
using two damage indices available in the literature.
A new phenomenological soil-structure interaction model is implemented in OpenSees
for analyses performed in the third part of this thesis. Soil variability is explicitly
considered and imposed on the numerical models to capture the random field variability
of the soil properties at site. The impact of the soil variability at the site is assessed at
the serviceability and ultimate limit states, and seismic fragility curves are developed
for the case that accounts for this source of soil parameter uncertainty.
Results in manuscript 1 indicate that for the vintage plan-asymmetric building
analyzed, the ground motion duration plays an important role in the damage estimation.
For the bridge models, in manuscript 2, the proposed retrofit solution applied to the
bridges columns provides a significant increase to the structural strength and
deformation capacity; however, the retrofit solution is not as efficient when the bridge
has columns with different lengths along its longitudinal direction versus the ideal case
when the bridge has columns with the same lengths. The ground motion duration plays
a role in the damage predicted by the two damage indices in both non-retrofitted and
retrofitted bridges.
Lastly, in manuscript 3, the inclusion of SSI effects increases the steel moment resisting
frame building measured response of peak and residual deformations, and the ground
motion duration plays an important role in the damage predicted when considering SSI
effects.
©Copyright by Andre Belejo
September 14, 2017
All Rights Reserved
Evaluation of ground motion duration effects on the damage prediction of building
and bridge structural and soil-structural systems

by
Andre Belejo

A THESIS

submitted to

Oregon State University

in partial fulfillment of
the requirements for the
degree of

Doctor of Philosophy

Presented September 14, 2017


Commencement June 2018
Doctor of Philosophy thesis of Andre Belejo presented on September 14, 2017

APPROVED:

Major Professor, representing Civil Engineering

Head of the School of Civil and Construction Engineering

Dean of the Graduate School

I understand that my thesis will become part of the permanent collection of Oregon
State University libraries. My signature below authorizes release of my thesis to any
reader upon request.

Andre Belejo, Author


ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to some people that directly or indirectly
contributed to make the elaboration of thesis possible.

To Dr. Andre Barbosa who was the main responsible for this achievement. He trusted
me. He received me like I was a family member. He made everything he could to make
me feel comfortable on my first time abroad. He provided all the necessary guidance,
insight and support to all the work performed. And more than an advisor he was a
friend. I will always consider him as reference on my career and personal life.

I also want to acknowledge Dr. Rita Bento, who also trusted me and made possible this
opportunity. She also gave important insight and support to the execution of this work.

I want to thank the contribution of Dr. Chris Higgins, Dr. Armin Stuedlein, Jonathan
Huffman, and Filipe Ribeiro for their insights, comments and discussions that helped
to build the foundations of this work.

To the committee members Dr. Armin Stuedlein, Dr. Michael Scott, Dr. Thomas
Miller, Dr. Yue Zhang and Dr. Julie Taylor for their kindness and availability. Not
forgetting all the helpful comments and discussions during our meetings and
examinations along this path.

To my parents, the ones I owe everything in this life. A special Thank you, for all your
support and advice you give me every day.

To my closest family and friends for all support and encouragement.


Last but not least, I would like to thank my beloved wife Lina Thomas. Her love,
encouragement and support are always present. She has been constantly helping me,
doing what she can for me, making my life much easier. I am not sure if I would be
able to finish this stage of my life without her.
CONTRIBUTION OF AUTHORS

Dr. Barbosa reviewed and made revisions to the main text of all manuscripts. He
contributed with helpful discussions and guidance on the execution of all manuscripts.

Dr. Rita Bento, Dr. Chris Higgins, and Dr. Armin Stuedlein reviewed and made
revisions to the main text of the first, second, and third manuscripts, respectively. In
addition, discussions with all of them regarding the results obtained and the main
content to be shown in each manuscript were extremely helpful.

Jonathan Huffman worked on the SSI model parameters generation used on the third
manuscript. He was also involved on important discussions.
TABLE OF CONTENTS

Page

1 GENERAL INTRODUCTION ...................................................................................1

1.1 Problem Statement .............................................................................................1

1.2 Purpose and Scope of this Dissertation ..............................................................2

1.3 Research Goal and Objectives ...........................................................................2

1.4 Organization of the Dissertation ........................................................................5

2 LITERATURE REVIEW ...........................................................................................7

2.1 Ground Motion Duration ...................................................................................7

2.2 Seismic Assessment of Irregular Building Structures .......................................9

2.3 Seismic Assessment of Bridges .......................................................................10

2.4 Retrofitting of non-seismically designed reinforced bridge columns ..............11

2.4.1 Laboratory testing ...................................................................................12

2.4.2 Lap-splice Modeling ...............................................................................17

2.5 Soil-Structure Interaction of Shallow Foundations Structure Systems............18

2.6 Damage Indices ................................................................................................22

2.7 Unresolved questions on the topic ...................................................................23

2.8 Bibliography ....................................................................................................25

3 INFLUENCE OF GROUND MOTION DURATION ON DAMAGE INDEX-


BASED FRAGILITY ASSESSMENT OF A PLAN-ASYMMETRIC NON-
DUCTILE REINFORCED CONCRETE BUILDING ................................................35
TABLE OF CONTENTS (Continued)

Page

3.1 Abstract ............................................................................................................36

3.2 Introduction ......................................................................................................36

3.3 Damage Indices ................................................................................................41

3.4 Ground motion selection ..................................................................................45

3.4.1 Long-duration ground motion set ...........................................................46

3.4.2 Short-duration ground motion set ...........................................................46

3.4.3 Ground Motions Characterization...........................................................48

3.5 Building case study ..........................................................................................53

3.5.1 General Description ................................................................................53

3.6 Nonlinear Finite Element Model Development and analysis ..........................55

3.6.1 Model Validation ....................................................................................56

3.7 Results and Discussion ....................................................................................60

3.7.1 Global response .......................................................................................60

3.7.2 Damage Assessment ...............................................................................65

3.7.3 Fragility Curve Development .................................................................71

3.8 Conclusions ......................................................................................................76

3.9 Acknowledgements ..........................................................................................79

3.10 References ......................................................................................................79


TABLE OF CONTENTS (Continued)

Page

4 DAMAGE-BASED FRAGILITY ASSESSMENT OF RC BRIDGES


RETROFITTED WITH TITANIUM ALLOY BARS ACCOUNTING FOR THE
EFFECTS OF GROUND MOTION DURATION ......................................................86

4.1 Abstract ............................................................................................................87

4.2 Introduction ......................................................................................................87

4.3 Methodology ....................................................................................................93

4.4 Case Studies .....................................................................................................94

4.5 Model Development and Validation ................................................................98

4.5.1 ... Column Modeling Considering a Modified Lap-splice Bond Stress-slip


Model for Use in Fiber Sections ..................................................................................98

4.5.2 TiAB Retrofit Modeling .......................................................................103

4.5.3 Combined Lap-splice and TiAB Retrofit Modeling .............................105

4.5.4 Bridge Models .......................................................................................106

4.5.5 Ground Motion Selection ......................................................................107

4.6 Results ............................................................................................................109

4.6.1 Modal Results .......................................................................................110

4.6.2 Time-history Response Analysis ..........................................................111

4.6.3 Damage assessment based on incremental dynamic analysis ...............120

4.7 Discussion ......................................................................................................125


TABLE OF CONTENTS (Continued)

Page

4.7.1 Effect of bridge retrofits using TiABs ..................................................125

4.7.2 Effect of bridge irregularity ..................................................................128

4.7.3 Ground motion duration effects ............................................................129

4.8 Conclusions ....................................................................................................129

4.9 References ......................................................................................................132

5 INFLUENCE OF GROUND-MOTION DURATION ON DAMAGE


ESTIMATION OF STEEL MOMENT RESISTING FRAMES ACCOUNTING FOR
SOIL-STRUCTURE INTERACTION EFFECTS ....................................................138

5.1 Overview ........................................................................................................139

5.2 Introduction ....................................................................................................140

5.3 Model Development.......................................................................................143

5.3.1 Soil Models ...........................................................................................143

5.3.2 Development of the nonlinear SSI model .............................................144

5.4 Building Models.............................................................................................148

5.5 Number of springs to assign to the bearing pressure model ..........................150

5.6 Methodology ..................................................................................................153

5.6.1 Foundation Design ................................................................................153

5.6.2 Nonlinear SSI model parameter generation ..........................................154

5.6.3 Structural Analysis ................................................................................157


TABLE OF CONTENTS (Continued)

Page

5.7 Results and discussion ...................................................................................159

5.7.1 Modal Analysis .....................................................................................159

5.7.2 Serviceability analysis ..........................................................................161

5.7.3 Preliminary results and damage assessment calibration .......................163

5.7.4 Influence of SSI on duration effects .....................................................167

5.8 Conclusions ....................................................................................................176

5.9 References ......................................................................................................178

6 CONCLUSIONS.....................................................................................................184

6.1 Summary of the Research Performed .............................................................184

6.2 Main Contributions..........................................................................................186

6.3 Overview of Main Findings ...........................................................................187

6.4 Future Work ....................................................................................................188

Appendices .................................................................................................................189
LIST OF FIGURES

Figure Page

Figure 1.1 – Flow-chart with overview of the work in this Dissertation and Links to
other work being performed by the research group that this work is tied to. ................4

Figure 2.1 – Design of the columns tested in the Oregon State University Structural
laboratory (Lostra 2016): Column C1 (control specimen), Column C2NL (specimen
without lap-splice), and column C2 (retrofitted specimen) .........................................18

Figure 2.2 – Construction of: (a) Control specimen; (b) Retrofitted specimen (Lostra
2016) ............................................................................................................................19

Figure 3.1 - Relation between arias intensity Ia and significant duration D5-75 of one
ground motion record ...................................................................................................40

Figure 3.2 – Response spectra for earthquake record pair #1, including the Tohoku
(2011) Towadako long-duration earthquake record and the scaled and rotated Hector
Mine (1999) Hector short-duration earthquake record: (a) Horizontal component H1;
(b) Horizontal component H2, and (c) Geometric mean of the two horizontal
components ..................................................................................................................49

Figure 3.3 – Acceleration time-series for earthquake records of pair #1: (a) Tohoku
(2011) Towadako station, horizontal component H1; (b) Hector Mine (1999) Hector
station, horizontal component H1; (c) Tohoku (2011) Towadako station, horizontal
component H2; and (d) Hector Mine (1999) Hector station, horizontal component
H2.................................................................................................................................50

Figure 3.4 – Response spectra for the ground motion set selected: (a) Horizontal
component H1; (b) Horizontal component H2 ..............................................................51
LIST OF FIGURES (Continued)

Figure Page

Figure 3.5 – Histograms for ground motions components for all earthquake records
used in this study: (a) Significant duration D5-75; (b) ratio of D5-75 of long-duration
motions to D5-75 of short-duration motions; and (c) scale factor applied to short-
duration earthquake records. ........................................................................................53

Figure 3.6 – Correlation coefficient between significant duration, D5-75, and natural
logarithm of the spectral acceleration versus period of vibration. ...............................53

Figure 3.7 – SPEAR building specimen (Fardis and Negro 2005) ..............................54

Figure 3.8 – Geometry and detailing of the Spear 3-story building: (a) Plan view; (b)
Elevation view; (c) Column cross sections; (d) Typical interior beam cross-section.
All dimensions are in meters........................................................................................55

Figure 3.9 – Correlation between the experimental and analytical results: (a) Roof
displacement in X direction at column C1; (b) Roof displacement in Y direction at
column C9; (c) Interstory drift in X direction at column C1 – 1st story; (d) Interstory
drift in X direction at column C9 - 2nd story. Experiment and analytical model
developed using OpenSees were subjected to the Hercegnovi record from the
Montenegro (1979) earthquake scaled to a PGA of 0.2g. ............................................59

Figure 3.10 – Observed damage on members after experiment (Negro et al. 2004)
with computed damage indices from numerical analysis: (a) Damage state of base of
column C3 - Minor damage; (b) Damage state of top of column C1 at 2nd story -
Moderate damage; (c) Damage state of top of column C3 at 2nd story - Severe
damage .........................................................................................................................60

Figure 3.11 – Incremental dynamic analysis curves for peak of the vector sum of
interstory drift ratios in X- and Y-directions. ..............................................................61
LIST OF FIGURES (Continued)

Figure Page

Figure 3.12 – Median interstory drift ratio envelopes: (a) C8 in X direction for,
considering all cases; (b) C2 in Y direction for Sa(T1) = 0.2g, considering all cases;
(c) C3 Vector Sum for Sa(T1) = 0.2g, considering all cases; d) C8 in X direction for
Sa(T1) = 0.4g, considering all cases; (e) C2 in Y direction for Sa(T1) = 0.4g,
considering all cases; (f) C3 Vector Sum for Sa(T1) = 0.4g, considering all cases; (g)
C8 in X direction for, considering only the non-collapsed cases; (h) C2 in Y direction
for Sa(T1) = 0.4g, considering only the non-collapsed cases and; (i) C3 Vector Sum
for Sa(T1) = 0.4g, considering only the non-collapsed cases .......................................63

Figure 3.13 – Statistics of normalized absolute roof displacements u at three column


locations (C8, C3, C2) relative to the absolute roof displacement of the centre of
stiffness uCR: (a) instant when peak u/uCR occurs at C2; (b) instant when peak u/uCR
occurs at C8..................................................................................................................64

Figure 3.14 – Sum of yielding energy dissipated in beam and column hinges for
short- and long-duration ground motion sets. ..............................................................66

Figure 3.15 – Park and Ang damage index results for structural alignment C5-C1-C2:
(a) Median damage values for columns at five intensities of shaking (0.1g, 0.2g, 0.3g,
0.4g, and 0.5g); (b) Median damage values for beams at five intensities of shaking
(0.1g, 0.2g, 0.3g, 0.4g, and 0.5g); (c) Damage index for hinges at column ends for the
Sa(T1) = 0.5g intensity level. ........................................................................................68

Figure 3.16 – Reinhorn and Valles damage index results for structural alignment C5-
C1-C2: (a) Median damage values for columns at five intensities of shaking (0.1g,
0.2g, 0.3g, 0.4g, and 0.5g); (b) Median damage values for beams at five intensities of
shaking (0.1g, 0.2g, 0.3g, 0.4g and 0.5g); (c) Damage index for hinges at column ends
for the Sa(T1) = 0.5g intensity level .............................................................................69
LIST OF FIGURES (Continued)

Figure Page

Figure 3.17 – (a) Global median DIP&A and story median DIP&A; (b) Global median
DIR&V and story median DIR&V ....................................................................................70

Figure 3.18 – (a) Envelope for median DIP&A; (b) Envelope for median DIR&V .........71

Figure 3.19 – Story and global structure fragility curves: (a) Moderate damage state
based on use of Park and Ang damage index; (b) Severe damage state based on use of
Park and Ang damage index; (c) Collapse based on use of Park and Ang index; (d)
Moderate damage state based on use of Reinhorn and Valles damage index; (e)
Severe damage state based on use of Reinhorn and Valles damage index; (f) Collapse
based on use of Reinhorn and Valles index .................................................................73

Figure 3.20 – Fragility curves based in peak interstory drift ratio IDR and on damage
indices DIP&A and DIR&V: (a) Probability of severe damage on the building subjected
to short-duration ground motions; (b) Probability of severe damage on the building
subjected to long-duration ground motions; (c) Probability of collapse of the building
subjected to short-duration ground motions; (d) Probability of collapse of the building
subjected to long-duration ground motions .................................................................74

Figure 4.1 – Columns tests reported in Lostra (2016): (a) Column C1 (reference
column), (b) Column C2, (c) Column C3, (d) Column C2NL, (e) typical cross-section
for column C1, and (f) typical cross section for columns C2 to C2NL in retrofitted
region. ..........................................................................................................................92

Figure 4.2 - Bridge geometry considered for the study cases: (a) Bridge RB1/RB1R
elevation view; (b) Bridges RB2/RB2R, IB1/IB1R, IB2/IB2R and IB3/IB3R elevation
view; (c) Bent 2 - Section DD’; (d) Deck reinforcement; (e) Cap beam reinforcement,
(f) Column and retrofit lengths assumed for all columns. ...........................................97
LIST OF FIGURES (Continued)

Figure Page

Figure 4.3 - (a) Bond stress – slip monotonic relationship (Harajli et al. 2004); (b)
Splitting of concrete during a laboratory test (Lostra 2016), as an example of the
perimeter of the characteristic block typically considered for the splitting failure
mechanism; (c) Scheme for the definition of the perimeter of the characteristic block
considered for the splitting failure mechanism; (d) Assumed stress distribution along
steel bars.....................................................................................................................101

Figure 4.4 – Validation of component response. Results measured in the laboratory


experiments and in the analytical model: (a) Column C1 load-displacement cyclic
response; (b) Column C1 – hysteretic energy dissipated; (c) Column C2NL load-
displacement cyclic response; (d) ) Response of TiABs to one load cycle; (e) Column
C2 load-displacement cyclic response ; (f) Column C2 – hysteretic energy
dissipated....................................................................................................................105

Figure 4.5 – Conceptual representation (a) Nonlinear finite element model developed
in Opensees; (b) Column model with TiAB retrofit. .................................................107

Figure 4.6 - (a) Components #1 of the response spectra of short- and long-duration
grounds motions of pair #6 (Table A4); (b) Components #2 of the response spectra
short- and long-duration grounds motions of pair #6 (Table A4); (c) Geometric Mean
of the response spectra of short- and long-duration grounds motions of pair #6 (Table
A4); d) Accelerograms of the components #1 of pair #6 (Table A4); (e) Correlation
coefficient between D5-75, and natural logarithm of the Sa(T1) versus period of
vibration, considering the ground motions selected for Bridge RB1/RB1R; (f)
Correlation coefficient between D5-75, and natural logarithm of Sa(T1) versus period
of vibration considering the ground motions selected for Bridge RB2/RB2R,
IB1/IB1R, IB2/IB2R, and IB3/IB3R .........................................................................109
LIST OF FIGURES (Continued)

Figure Page

Figure 4.7 - First mode shapes and respective periods: (a) Bridge RB1/RB1R; (b)
Bridge RB2/RB2R; (c) Bridge IB1/IB1R; (d) Bridge IB2/IB2R; (e) Bridge
IB3/IB3R ....................................................................................................................111

Figure 4.8 – Time-history of drift ratios of the non-retrofitted -and retrofitted column
C1-1: (a) Bridge RB1/RB1R along the longitudinal direction; (b) Bridge RB1/RB1R
along the transverse direction; (c) Bridge RB2/RB2R along the longitudinal direction;
(d) Bridge RB2/RB2R along the transverse direction; (e) Bridge IB1/IB1R along the
longitudinal direction; and (f) Bridge IB1/IB1R along the transverse direction; ......112

Figure 4.9 –Energy dissipated at the top column along between non-retrofitted -and
retrofitted column C1-1: (a) Bridge RB1/RB1R along longitudinal direction; (b)
Bridge RB1/RB1R along transverse direction; (c) Bridge RB2/RB2R along
longitudinal direction; (d) Bridge RB2/RB2R along transverse direction; (e) Bridge
IB1/IB1R along longitudinal direction; and (f) Bridge IB1/IB1R along transverse
direction;. ...................................................................................................................113

Figure 4.10 - Energy Dissipated in the column top along longitudinal direction: (a)
Different positions of short “1” and median length columns “2” (Bridges IB1, IB2
and IB3) for Sa(T1) = 0.4g; (b) Different positions of short “1” and median length
columns “2” (Bridges IB1R, IB2R and IB3R) for Sa(T1) = 0.8g; (c) Comparison
between short and long direction in Bridge RB1; (d) Comparison between short- and
long-duration in Bridges RB2 and IB1. .....................................................................114

Figure 4.11 - Column shear vs drift ratio time-history response along the longitudinal
direction of bridge IB1: (a) column C1-1 (short column); (b) column C3-1 (long
column) ......................................................................................................................115
LIST OF FIGURES (Continued)

Figure Page

Figure 4.12 - Surface plots that relate D5-75, Sa(T1) Peak drift ratio: (a) Bridge RB1;
(b) Bridge RB2 (c) Bridge IB1; (d) Bridge RB1R; (e) Bridge RB2R; (f) Bridge
IB1R ...........................................................................................................................117

Figure 4.13 - Surface plots that relate D5-75, Sa(T1) Damage Index Park and Ang: (a)
Bridge RB1; (b) Bridge RB2 (c) Bridge IB1; (d) Most damaged element of Bridge
IB1; (e) Bridge RB1R; (f) Bridge RB2R (g) Bridge IB1R; (h) Most damaged element
of Bridge IB1R ...........................................................................................................118

Figure 4.14 - Surface plots that relate D5-75, Sa(T1) and Damage Index Reinhorn and
Valles: (a) Bridge RB1; (b) Bridge RB2 (c) Bridge IB1; (d) Most damaged element of
Bridge IB1; (e) Bridge RB1R; (f) Bridge RB2R (g) Bridge IB1R; (h) Most damaged
element of Bridge IB1R .............................................................................................119

Figure 4.15 - Median incremental dynamic analysis curves for vector sum of peak
drift ratios in X- and Y-directions based on results from bridge models: (a) RB1 and
RB1R; (b) RB2 and RB2R; and (c) IB1 and IB1R. ...................................................121

Figure 4.16 – Median incremental dynamic analysis damage curves: (a) DIP&A curves
for RB1 and RB1R; (b) DIP&A curves for RB2 and RB2R; (c) DIP&A curves for IB1
and IB1R; (d) DIR&V curves for RB1 and RB1R; (e) DIR&V curves for RB2 and
RB2R; and (f) DIR&V curves for IB1 and IB1R. ........................................................122
LIST OF FIGURES (Continued)

Figure Page

Figure 4.17 - (a) Median IDA curves computing the peak vector sum of drift ratios in
X- and Y-directions; (b) Evolution of median DIP&A curves with the increase of
seismic intensities; (c) Evolution of median DIR&V curves with the increase of seismic
intensities; (d) Fragility curves that compute probability of exceeding a determined
drift ratio (4% for the non-retrofitted bridges and 10% for the retrofitted bridges); (e)
Fragility curves that compute probability of reaching collapse (DI > 0.8) based on the
Park and Ang damage index DIP&A; and (f) Fragility curves that compute probability
of reaching collapse (DI>0.8) based on the Reinhorn and Valles damage index
DIR&V..........................................................................................................................124

Figure 5.1– Example of spatial variability of undrained shear strength (su,mean = 70


kPa, COV = 15% and δh = 10 m). ..............................................................................144

Figure 5.2 – Bearing pressure-displacement curves used to fit the hyperbolic model
developed in Huffman et al. (2015). ..........................................................................145

Figure 5.3–(a) Bearing pressure-displacement foundation model implemented in


OpenSees as the HyperbolicQy material model; (b) HypergolicGap model used to
mobilize the passive resistance of the footing (Duncan and Mokwa 2001); and (c)
Modified TzSimple2 model to capture the frictional component of load and
displacement along base of the footing. .....................................................................148

Figure 5.4 – (a) Two-dimensional models of steel moment resisting frames buildings;
(b) Identification of all column and beams sections; (c) Column (fiber section) and
beam (linear section) sections. ...................................................................................150

Figure 5.5 - General model of the SSI model implemented. .....................................150


LIST OF FIGURES (Continued)

Figure Page

Figure 5.6 - Deviations obtained (from analysis with 101 springs in each footing) for
the calibration of the number of springs to use at each footing of the 3-story building:
(a) Angular distortion, (b) Interstory drift ratio, (c) Peak beam end-moments..........152

Figure 5.7 – Deviations obtained (from analysis with 101 springs in each footing) for
the calibration of the number of springs to use at each footing of the 9-story building:
(a) Angular distortion, (b) Interstory drift ratio, (c) Peak beam end-moments..........153

Figure 5.8 - Copulas showing the parameters correlations used for footing #1 of the 3-
story building considering 50,000 parameter sets, with variability associated to
COV=15% and δh = 10 m. Points in red correspond to the reduced data set : (a) k1-k2,
(b) k1-qSTC, and (c) k2-qSTC; (d) k1-1/k2; (e) 1/k1-qSTC ...............................................157

Figure 5.9 – First three periods of vibration of the building models with different
scenarios normalized by the fundamental period of the fixed-base structure, and the
SSI models of the: (a) 3-story building, (b) 9-story building. ...................................160

Figure 5.10 - Cumulative density functions (CDF) of the angular distortion values
obtained for the 3-story building, considering the “allowable”, service and ultimate
limit states: (a) each bay, considering all variability cases, (b) each variability type for
the left bay..................................................................................................................161

Figure 5.11 - Cumulative density functions (CDF) of the angular distortion values
obtained for the 9-story building considering the “allowable”, service and ultimate
limit states: (a) each bay, considering all variability cases, (b) each variability type for
the left bay..................................................................................................................162
LIST OF FIGURES (Continued)

Figure Page

Figure 5.12 – Variation of the deformation demands and global damage on the
buildings with increase of fundamental period of the soil-structure system: (a) 3-story
building: ratio of mean peak IDR considering SSI with fixed-base model peak IDR;
(b) 3-story building: ratio of mean DIP&A considering SSI with fixed-base model
DIP&A; (c) 9-story building: ratio of mean peak IDR considering SSI with fixed-base
model peak IDR; (d) 9-story building: ratio of mean DIP&A considering SSI with
fixed-base model DIP&A; ............................................................................................165

Figure 5.13 – 3-story building responses and surface response functions of the
significant duration D5-75 and spectral acceleration at the fundamental period of the
fixed-base model Sa(T1Fixed-base) for the 44 unscaled ground motion and the 80 SSI
different cases (when considered): (a) Fixed-base model global DIP&A, (b) Fixed-base
model global DIR&V, (c) Fixed-base model peak IDR, (d) SSI model global DIP&A, (e)
SSI model global DIR&V, (f) SSI model peak IDR.....................................................166

Figure 5.14 - 9-story building response function of the significant duration D5-75 and
spectral acceleration at the fundamental period of the fixed-base model Sa(T1Fixed-base)
for the 44 unscaled ground motion and the 80 SSI different cases (when considered):
(a) Fixed-base model global DIP&A , (b) Fixed-base model global DIR&V, (c) Fixed-
base model peak IDR, (d) SSI model global DIP&A, (e) SSI model global DIR&V, (f)
SSI model peak IDR. .................................................................................................167

Figure 5.15 – History of IDRs during one short-duration and one long-duration
ground motions considered both fixed base and SSI models: (a) 3-story building 1st
and 3rd stories; (b) 9-story building 2nd and 8th stories. ..............................................169
LIST OF FIGURES (Continued)

Figure Page

Figure 5.16 – Energy dissipated by yielding of the beams not considering SSI effects,
energy dissipated by yielding of beams considering SSI effects, and vertical and
horizontal springs that compose the SSI model: (a) 3-story building; and (b) 9-story
building. .....................................................................................................................170

Figure 5.17 – IDA curves comparing median responses of fixed-base and SSI models
of the 3-story building subjected to short- and long-duration ground motions in terms
of: (a) peak interstory drift ratios, and (b) Residual interstory drift ratios ................172

Figure 5.18 – IDA curves comparing median responses of fixed-base and SSI models
of the 9-story building subjected to short- and long-duration ground motions in terms
of: (a) peak interstory drift ratios, and (b) Residual interstory drift ratios ................172

Figure 5.19 - Curves comparing median damage indices of the fixed-base and SSI
models of the 3-story building subjected to short- and long-duration ground (a)
DIP&AMax.member, and (b) global DIP&A, (c) DIR&VMax.member, and (d) global DIR&V .....174

Figure 5.20 - Curves comparing median damage indices of the fixed-base and SSI
models of the 9-story building subjected to short- and long-duration ground (a)
DIP&AMax.member, and (b) global DIP&A, (c) DIR&VMax.member, and (d) global DIR&V .....175
LIST OF TABLES

Table Page

Table 3.1 – Damage Index Classification ....................................................................45

Table 3.2 – Numerical model and experiment specimen modal properties of the 3-
Story building...............................................................................................................57

Table 4.1 - Reduction of probability of reaching a specific damage state based on


DIP&A and DIR&V for a ground motion intensity of Sa(T1) = 0.5g ..............................126

Table 4.2 - Reduction of probability of reaching a specific damage state based on


DIP&A and DIR&V for a ground motion intensity of Sa(T1) = 1.0g ..............................127

Table 4.3 - Reduction of probability of reaching a specific damage state based on


DIP&A and DIR&V for a ground motion intensity of Sa(T1) = 1.5g ..............................128

Table 5.1 – Footings dimensions of buildings studied ..............................................154


1

1 GENERAL INTRODUCTION

1.1 Problem Statement

Recent great earthquakes such as the Chile, Maule 2010 (Mw = 8.8) or Tohoku, Japan 2011
(Mw = 9.0) caused significant structural damage to the built environment. The ground motions
recorded in seismic stations during these events are characterized by having significantly longer
duration than ground motions typically recorded from major crustal earthquakes (Mw = 7.0 to
7.9) in other locations in the World. Other events such as the 1700 Cascadia Subduction Zone
Great Earthquake (~M 9.0) in the United States, and the great Lisbon earthquake of November
1st, 1755 (~M 8.7) in western Europe were megathrust earthquakes assumed to have produced
similar long ground motion durations.

While it could be intuitive to say that longer duration of shaking would results in increased
damage due to these strong shaking events, previous studies on the effect of earthquake ground
motion duration on the assessment of civil infra-structures have provided mixed results, with
some indicating a clear increase in damage and peak deformation responses, while others
showing limited effect of the increased ground motion duration on the seismic response or even
showing reduced residual post-earthquake drifts. In part, these mixed results can be attributed to
the lack of detailed structural modeling that captures cyclic degradation of structural components
of building and bridge structures when subjected to earthquakes and in part due to the lack of
ground motion records from the large magnitude long-duration records. Other reasons may exist,
and in some cases it is reasonable to expect that the ground motion duration may not play a role,
especially for reduced intensities of shaking.

Based on existing knowledge and on new ground motion datasets recorded in the recent
decade, this dissertation provides a summary of the literature available on the effects of ground
motion duration on performance of civil infra-structures, and describes the work developed by
the author on the effect of ground motion duration on damage and fragility assessment of
building and bridge structures. The assessment of existing and retrofitted structures,
2

consideration of soil-structure interaction effects, and the study of structural irregularity are
topics that are addressed in this work.

1.2 Purpose and Scope of this Dissertation

There is a lack of consensus in the existing literature regarding the effects of ground motion
duration. Thus, ground motion duration is not considered by the current seismic codes for design
of new structures or assessment of existing structures. However, recent studies indicate that
ground motion duration does affect the seismic behavior of structures.

This document provides new knowledge to improve the understanding of earthquake ground
motion duration effects on buildings and bridges. In particular, the effect of ground motion on
the seismic assessment is studied in combination with the following subtopics: (1) structural
irregularity, which is characterized by plan-asymmetry in buildings, or by different column
heights in bridges, (2) soil-structure interaction (SSI), and (3) structural seismic retrofit.

1.3 Research Goal and Objectives

The main goal of this study is to improve the understanding of the role of ground motion
duration on the seismic performance of civil infra-structures. In order to accomplish this goal,
more specific objectives of this work are to:

 Develop an understanding of the effects of ground motion duration on the response of


plan-asymmetric reinforced concrete (RC) building structures when subjected to long-
duration and short-duration earthquake ground motions;
 Understand the effects of ground motion duration on the response of existing bridges
with regular and irregular geometry;
 Develop a modeling strategy to capture the experimentally observed response of existing
columns and columns retrofitted using titanium alloy bars (TiABs);
 Characterize the uncertainty of the response of bridges when subjected to short-duration
and long-duration earthquake ground motions, including those retrofitted TiABs;
3

 Gain an understanding of how the combined effects of ground motion duration and soil-
structure interaction impact the structural performance of buildings, focusing on pre-
Northridge moment resisting steel frame buildings as application examples.
Evaluation of ground motion duration
effects on the damage prediction of
building and bridge structural and soil-
structural systems

Buildings Bridges

Retrofitted Vintage Steel Vintage RC Vintage RC Retrofitted


Steel Buildings Buildings Buildings Bridges RC Bridges

Ribeiro et al. (201X) Barbosa et al. 2017

Effect of Effect of Effect of


Duration? Duration? Duration?

Manuscript 1 Manuscript 2
Belejo et al. (2017) Belejo et al. (201X)
Published in Journal of (Prepared for Submission on Journal
Engineering Structures of Bridge Engineering ASCE)
Soil-Structure
Interaction for
Serviceability

work being performed by the research group that this work is tied to.
Huffman et al. (2015) Manuscript 3
Effect Belejo et al. (201X)
of SSI? (Prepared for Submission on Journal of
Structural Engineering ASCE)

Figure 1.1 – Flow-chart with overview of the work in this Dissertation and Links to other
4
5

1.4 Organization of the Dissertation

This dissertation starts with this introductory chapter that describes the motivation to work
on the topic, the research objectives to be accomplished, and the methodology used to tackle
these objectives.

The format of the document follows a manuscript format, and it includes three main
manuscripts that were developed as part of the work as shown in Figure 1.1.

Chapter 2 presents the literature review on the studies related with topics that are
fundamental for the three manuscripts, including: (1) seismic assessment of irregular structures;
(2) recent work performed on assessment of bridges; (3) laboratory tests using titanium alloy
bars (TiABs); (4) assessment of SSI effects on the structural behavior of buildings or bridges; (5)
seismic damage indices, proposed in order to quantify and qualify structural damage; and (6)
assessment of ground motion duration effects.

Chapter 3 presents the work developed on the assessment of ground motion effects on a 3-
story RC building that was tested in Europe, which is also referred to as Manuscript 1. In this
chapter, a three-dimensional nonlinear model of a full-scale building that was tested in the ELSA
laboratories in Europe is validated for studying the ground motion duration effects. The building
model is subjected to two ground motion sets: (1) long-duration bi-directional earthquake ground
motions, and (2) short-duration bi-directional earthquake ground motions that were selected
through a new selection procedure for assessment of duration effects on 3-D structures. A
damage assessment is performed analyzing global and local responses. A methodology for
development of new damage index-based fragility functions is proposed in this study.

Chapter 4 presents the work developed for Manuscript 2. This chapter focuses on the
assessment of ground motion effects on typical reinforced concrete bridges constructed between
1950 and 1970 in the State of Oregon. Nonlinear models of as-built and retrofitted bridges with
titanium alloy bars are developed and validated using results obtained from laboratory testing
performed at Oregon State University (Lostra 2016). The nonlinear finite element models are
subjected to short- and long-duration ground motion sets, using the procedures developed in
6

Chapter 3 for ground motion selection for assessment of ground motion duration on the response
of structures. Damage assessment of the bridge models is performed analyzing global and local
bridge responses. In this chapter, the impact of the retrofit solution, ground motion duration
effects, and bridge geometry irregularity are studied.

Chapter 5, which corresponds to Manuscript 3, summarizes work performed on the


assessment of the effects of ground motion duration of strong shaking on the seismic response of
two steel moment frame buildings considering soil-structure interaction (SSI). The building
structural response is evaluated considering the SSI effects by implementing a new
phenomenological spring model that is calibrated based on a large dataset collected by
colleagues at Oregon State University. Uncertainty in soil parameters and their spatial
distribution is explicitly considered in the analyses performed in this work. The impact of
duration previously identified on these buildings is addressed and compared with the impact
observed considering SSI.

Chapter 6 presents a summary of the main conclusions taken in this work, and proposes
challenges and recommendations for future work.

As part of the manuscript format used to present this thesis, references are included at the
end of each chapter according to the respective citations in the text presented for each
manuscript.
7

2 LITERATURE REVIEW

This chapter presents a review of literature where the main research topics studied in this
work are addressed, including: (1) effect of earthquake ground motion duration on seismic
response of building and bridge structures, (2) seismic assessment on irregular building
structures, (3) seismic fragility assessment of bridges, (4) retrofitting of bridge columns with
short lap-splices, including recent laboratory tests using TiABs, (5) analytical solutions to
account with SSI effects, (6) damage indices for quantification seismic damage.

2.1 Ground Motion Duration

Several studies over the past decades have researched the effect of ground motion duration
on structural response, often with inconclusive results (Nassar and Krawinkler, 1991; Shome et
al., 1998; Tremblay and Atkinson, 2001; Chai, 2005). Results in other studies (Cornell, 1997;
Bommer et al., 2004; Bojorquez et al., 2006; Hancock and Bommer, 2006; Iervolino et al., 2006)
indicated that there is no significant correlation between measures of significant duration
(duration of an interval across which a specified amount of energy is dissipated) and peak
interstory drifts (IDRs) in building structures. Nonetheless, in terms of the impact of duration on
peak IDRs, Chandramohan et al. (2016) presented results showing that long-duration ground
motions can induce higher deformations for large shaking intensities. Other results in studies that
focused on cumulative response parameters such as Oyarzo-Vera and Chow (2008), Dutta and
Mander (2001), and Manfredi et al. (2003) indicated that the amount of accumulated damage
depends on the duration. Ruiz-Garcia (2010) observed larger residual deformations induced by
long-duration ground motions. Foschaar et al. (2012), Raghunandan and Liel (2013), and Song et
al. (2014) highlighted that increased probability of collapse is related with the longer duration
ground motion. Raghunandan et al. (2015) performed a collapse assessment of structures
subjected to crustal and subduction zone ground motions, which typically exhibited longer
durations, and results indicated that there was a greater risk of structural collapse for subduction
zone ground motions. Results in Chandramohan et al. (2016) also indicate that higher probability
of structural collapse is expected when structures are subjected to long-duration motions.
Barbosa et al. (2017) quantified the influence of ground motion duration on the damage of a 3-,
9-, and 20-story steel building using two damage indices. Results in Barbosa et al. (2017)
8

indicate that the increase in energy dissipation demands in the structures subjected to long-
duration motions increased significantly the expected levels of structural damage at the higher
intensities of shaking.

The effect of duration in bridge structures has been addressed in a few recent studies.
Chandramohan et al. (2016) studied the effect of ground motion duration on a RC bridge pier
model applying two different sets of ground motions: a set of short- and a set of long-duration
ground motions. The median collapse capacity estimated when applying the set of long-duration
ground motions was inferior when compared to the median collapse capacity estimated by
applying the short-duration ground motion set. Dusicka and Lopez (2016) developed a model to
obtain the acceleration values expected from a full-rupture Cascadia Subduction Zone event. The
authors tested three circular laps-spliced reinforced concrete columns using crustal and
subduction ground motion. The results obtained indicate that ground motion duration can affect
the damage and deformation capacity of bridge columns. Bazaez and Dusicka (2016a) developed
quasi-static loading protocols that reflect greater inelastic demands of bridge columns subjected
to long-duration ground motions caused by subduction megathrust earthquakes, improving the
seismic assessment of bridge columns. In a companion paper (Bazaez and Dusicka 2016b) these
protocols were applied together with large scale experiments to evaluate a bridge pier retrofit
solution using buckling restrained braces (BRBs). The results showed that the retrofit solution
employed increases significantly the ductility of the retrofitted structure.

Ground motion duration effects have been taken into account on the assessment of the
behavior of the soil inducing greater soil deformations and probability of occurring liquefaction
of soil deposits. Generation of excess pore pressure have been observed in potentially liquefiable
soils, influenced by the amplitude, frequency content, duration, and phase characteristics of
earthquake ground motions (Kayen and Mitchell 1997; Kramer and Mitchell 2006; Sideras and
Kramer 2012).

It is worth noting also, that there are several measures that can be used to characterize the
ground motion duration for an earthquake record. In terms of quantification of long-duration
ground motion, Chandramohan et al. (2016) classified an earthquake record as “long-duration
earthquake” when at least one of its orthogonal ground motion components has a significant
9

duration (D5-75) greater than 25 seconds. Other definitions were also tested (e.g. D5-95, or
cumulative absolute velocity) and it was shown that D5-75 was a better predictor of the effects of
ground motion duration. The D5-75 ground motion measure was then used also in Barbosa et al.
(2017). This measure of ground motion duration is the measure of duration used in this
dissertation.

2.2 Seismic Assessment of Irregular Building Structures

Plan asymmetry of strength and stiffness in a building, also known as torsional irregularity,
is known to produce larger demands on some elements of the building, which in turn lead to
often premature and unexpected failures under seismic loading. For example, Bozorgnia and Tso
(1984) concluded that the effect of asymmetry on the element ductility demand and on edge
displacements plays an important role for stiff systems, reducing the yield strength of the
elements and consequently the building structure. Paulay (1997) described how the asymmetry in
mass, stiffness and strength distributions that cause torsional behavior in irregular buildings,
make it impossible to reproduce the seismic effects of irregular buildings using 2-D models,
indicating that 3-D building models should always be used. Stathopoulos and Anagnostopoulos
(2005) showed that response measures such as ductility factors and damage indices in the
flexible side of plan asymmetric buildings are up to two times greater than those at the stiff side
leading to premature member failures. Dutta and Das (2002), De-la-Colina (2003), Tena-
Colunga (2004), and Sommer and Bachmann (2005) performed seismic assessments of 3-D
structural models of code-designed plan irregular structures and concluded that demands in
irregular structures far exceeded the demands estimated using codified equations (e.g. EN 1998-
1, 2004; ICB, 1997). Faggella et al. (2013) illustrated the effects of scaling the time-series
accelerations of both components of the two orthogonal directions and the impact on the
structural demands of an irregular RC moment frame building structure.

Even though the previous paragraph highlights the impact of torsion irregularity in building
response, in the existing body of literature on the effects of ground motion duration on structural
response, only a very limited number of studies were identified that studied the effects of ground
motion duration using the three-dimensional (3-D) structural models of irregular structures (e.g.
Chakroborty and Roy, 2013). Roy and Chakroborty (2013) indicate that seismic vulnerability of
10

plan-asymmetric structures is sensitive to the characteristics of ground motions, including ground


motion duration. Nonetheless, since the study focused on multiple parameters a detailed analysis
of the effects of ground motion duration on the assessed damage was not performed.

2.3 Seismic Assessment of Bridges

Priestley et al. (1994) defined a methodology to perform seismic assessment of existing


bridges, divided into two stages: (1) screening and prioritization study to select the most
vulnerable bridges, and (2) detailed structural analysis considering the site seismicity and soil
conditions. Several topics were addressed, including: (1) different assessment procedures, (2)
limit states, (3) analysis schemes, (4) structural element strength, and (5) structural and soil
element deformation characteristics. The discussion of these topics was complemented with
recommendations on how to tackle them when performing a seismic bridge assessment. In
general, it could be said that the seismic fragility assessments that followed this work (Priestley
et al. 1994) make use of the general procedures of assessment summarized in Priestley et al.
(1994), using nonlinear static or nonlinear dynamic response analyses.

Mackie (2005) and Mackie and Stojadinović (2005, 2006) presented a methodology and
developed bridge fragility functions for highway overpass bridges in California. Zhang et al.
(2008) introduced a non-iterative method, based on the Pacific Earthquake Engineering Research
Center framework, for design of bridge structures in terms of single or multiple physical design
parameters, such as column height and diameter, considering uncertainty in the hazard, demand,
damage, and loss to the structure. Design equations are derived for different performance levels
that are defined in terms of bridge damage and loss. Gidaris et al. (2016) provides a state-of-the-
art on methods for fragility assessment of bridges to different hazards for existing bridges. The
work provides a general overview of probabilistic seismic and damage assessment developed
over the past couple of decades.

In terms of retrofitted bridges, Kim and Shinozouka (2004) presented a methodology for
developing and comparing the improvement in the fragility curves of bridges before and after
retrofit. To formulate the problem of fragility enhancement, the quantification was made by
comparing the median values of the fragility curves before and after the retrofit. Padgett and
DesRoches (2009) presented a methodology for developing fragility curves for retrofitted bridge
11

systems, where importance of evaluating the impact of the retrofit on the overall bridge fragility
is emphasized. The authors presented results on the influence of the different retrofit measures on
the fragility of each class of typical bridges in the Central and Southeastern United States.

In terms of assessment of bridges with irregular geometries based on the columns length,
there is an extensive number of references that address the topic available in the literature. For
example, Kappos et al. (2002) studied the effect of the soil-structure interaction (SSI) and pier
stiffness on the seismic response of RC bridges with irregular configurations concluding that
including SSI effects can be advantageous in the seismic design of irregular RC bridges, and that
small variations in the pier stiffness may lead to significant changes in the predominant modes of
vibration on irregular bridges. Pinho et al. (2009) applied commonly employed nonlinear static
procedures to regular and irregular bridge configurations to study the adequacy of these
procedures on the seismic assessment of existing continuous span bridges and concluded that all
methods are able to predict displacement response with good accuracy, while forces are well
captured only when using procedures that account for the higher modes of the bridges. Tamanani
et al. (2014) performed nonlinear pushover analyses to evaluate the adequacy of Eurocode 8 and
AASHTO (EN 1998-1, 2004; AASHTO, 2012) design procedures for seismic behavior of
irregular bridges, and investigated different ways of regularizing the bridge performance in order
to balance damage, obtaining a simultaneous or near-simultaneous failure of all unequal piers
independently of their height. The authors concluded that bridges with piers of unequal height
can develop a synchronized failure by having suitable arrangements of transverse reinforcement,
while the shortest piers have to be designed for higher ductility capacity.

2.4 Retrofitting of non-seismically designed reinforced bridge columns

Upon review of vintage Oregon bridge designs, the most common bridge types constructed
in the 1950s and 1960s used square reinforced concrete columns with insufficient flexural and
shear reinforcement to resist the expected seismic demands. In addition, the columns have poor
detailing of the lap-splices at the base of the column above the footing, creating a bond-slip
failure mode that significantly reduces the strength, stiffness, and displacement capacity of the
columns.
12

2.4.1 Laboratory testing


Several studies (Cairns and Arthur, 1979; Paulay, 1982; Lukose et al., 1982; Girard and
Bastien, 2002; Melek and Wallace, 2004; ElGawady et al., 2010) concluded that short lap-splice
lengths and inadequate transverse reinforcement are the main causes of bad performance of
vintage columns subjected to cyclic loads. For the case of short lap-splice lengths, they are
insufficient to fully develop the column reinforcement steel before the lap-splice bond strength is
exceeded, leading to a bond-slip failure mode in which the starter bars anchored in the footing
and the column bars slide relative to each other, reducing the strength, stiffness, displacement
capacity, and strength of the columns when subjected to cyclic lateral loading.

Lukose et al. (1982) tested ten (10) full-scale column specimens to study the behavior of
bond stresses and occurrence of bond failure from reversed cyclic loading. They concluded that
reversed cyclic loads were more detrimental to the performance of the splices compared to
monotonic loads. The same authors also studied the difference between bar slip in compression
and tension lap-splices and concluded that the contribution of bond stress was lower on the
tension splice when compared with the compression splice. Paulay (1982) suggested an increase
on the confinement along the lap-splice length of resisting columns could be a good solution to
improve the poor performance of the short lap-splices. Lynn et al. (1996) conducted a study on
columns with short lap-splices and widely spaced transverse hoops subjected to cyclic load
reversals, observing multiple failure mechanisms, including localized concrete crushing, rebar
buckling, lap-splice failure, and axial load collapse. In addition, vertical cracking occurred along
the lap-splice lengths and then cracking progressed up the column at higher drift displacements.
Seible et al. (1997) observed dilation and cover concrete spalling along the lap-splice length
when debonding occurs due to lap-splice failure in cyclic response of columns when poor lap-
splice detailing was combined with insufficient transverse reinforcing. Melek and Wallace
(2004) conducted a study on the cyclic behavior of reinforced concrete of full-scale columns
with short lap-splices subjected to axial and lateral load cycles, concluding that columns lose
their ductility capacity once lateral strength degradation from bond deterioration between the
reinforcement bars and surrounding concrete occurs. Nonetheless, the columns are able to
maintain their axial capacity to larger levels of drift, under displacement controlled testing.
13

When inadequate transverse reinforcement is an issue, buckling of the longitudinal


reinforcement bars occur between tie locations resulting in improperly confined concrete. The
compressive axial capacity is reduced and the concrete exhibits a brittle behavior. Geometry,
concrete strength, and transverse reinforcement have an important role to ensure an effective
confinement. A number of studies have addressed specific column parameters and their role in
confinement effectiveness. Sheikh and Uzumeri (1980) studied the relationship between
distribution of longitudinal steel and concrete confinement and ductility of square tied columns
and reached the conclusion that bars distributed evenly along the tie perimeter provided better
confinement than larger bars less evenly distributed. Moehle and Cavanagh (1985) assessed the
effectiveness of ties in concrete members for achieving confinement for ductile design.
Transverse reinforcement plays a fundamental role in maintaining core concrete strength in
columns subjected to inelastic load reversals by confining the concrete and improves the bond
between the longitudinal reinforcement and surrounding concrete. Paulay and Priestley (1992)
presented the compression stress-strain relationships from the Mander model for confined
concrete (Mander et al. 1988) for tied and continuously confined sections, demonstrating greater
efficiency of circular transverse reinforcement to achieve confinement in concrete columns
compared to rectangular ties. Watson et al. (1994) addressed the effects of concrete grade,
transverse steel configuration, and axial load in the ductility capacity of confined concrete
columns, finding that quantity of transverse reinforcement should be increased when the level of
axial load or the concrete strength increase, and can be decreased for greater longitudinal
reinforcement ratios.

Lostra et al. (2017) reported on laboratory tests the use of a novel solution for retrofitting
seismically vulnerable square reinforced concrete columns using externally mounted titanium
alloy bars (TiABs), which had previously only been tested on bridge girders under gravity
loading (Amneus 2015; Barker 2015). The use of TiABs expands the options available to
designers for improving the seismic performance of older reinforced concrete structures that do
not meet modern design requirements. Among the options available for bridge columns
retrofitting, the following have been the most commonly employed: (1) Carbon and Glass Fiber-
Reinforced Polymers (Katsumata et al., 1998; Feng et al., 1999; Iacobucci et al., 2003; Schlick
and Breña, 2004; Chang et al., 2004; Endeshaw et al., 2008; ElGawady et al., 2010); (2) Steel
14

Jackets (McLean and Bernards 1992; Abedi et al. 2010), and (3) Buckling Restrained Braces
(Bazaez and Dusicka 2016b).

Lostra et al. (2017) illustrated how the well-defined material properties of TiABs can
provide not only effective but also economical seismic strengthening of RC. TiABs have lower
stiffness but higher strength and ductility when compared to conventional reinforcing steel. In
addition, they are not affected by corrosion, thus having enhanced environmental durability,
which makes them a viable long-term solution for column strengthening.

The use TiABs for retrofitting RC columns is a solution that was explored by Lostra et al.
(2017) for vintage columns constructed using poor seismic detailing, namely insufficient length
for lap-splices of column bars immediately above the footing.

The retrofit solution using TiABs in Lostra et al. (2017) consisted of two parts: (1) vertical
TiABs were embedded into epoxy-filled drilled holes in the footings and columns with the
objective of increase the flexural capacity and develop a self-centering or restoring mechanism in
the column; and (2) concrete shell encased by a TiAB continuous spiral was added to provide
confinement to the column core and at the same time provide as bracing of the vertical TiABs. It
is worth noting that the TiABs were unbounded along their length, which allows for inspection
of the members.

The experimental results from three column tests presented by Lostra et al. (2017) include: a
control specimen (conventionally reinforced with detailing representative of vintage column
designs that did not consider seismic provisions), and two retrofitted specimens with externally
mounted TiABs acting as flexural ligaments and with a continuous TiAB spiral reinforced
confining shell. The conventionally reinforced columns had lap slices with a length of 0.90 m (3
feet) above the top of the footing. The two retrofitted specimens are: (1) a strengthened zone
extended 0.60 m (2-feet) above the lap-splice, and (2) a strengthened zone extending 0.40 m (1
foot-4 inches) above the lap-splice. A forth specimen with strengthened zone extended of 0.60 m
but without lap-splice was also tested. While the experimental testing program was performed
and results look promising, no analytical work was developed to characterize the performance of
other columns and bridge configurations not tested. The design features of the of the control
15

specimen, retrofitted specimen and specimen without lap-splice are shown in Figure 2.1. Figure
2.2 shows the control and retrofitted specimens during their construction process.
16

Figure 2.1 – Design of the columns tested in the Oregon State University Structural
laboratory (Lostra 2016): Column C1 (control specimen), Column C2NL (specimen without lap-
splice), and column C2 (retrofitted specimen)
17

(a) (b)

Figure 2.2 – Construction of: (a) Control specimen; (b) Retrofitted specimen (Lostra 2016)

2.4.2 Lap-splice Modeling


Inadequate lap-splice length is one of the most common issues in existing non-seismically
designed columns. Therefore, a numerical model able to capture the lap-splice of these columns
behavior is fundamental when simulating the lateral response of non-seismically designed
columns. Several modeling approaches that are available in the literature on the lap-splice
behavior are described next (Harajli et al., 2004; Zhao and Sritharan, 2007; Harajli, 2009;
Chowdhury and Orakcal, 2012).

Harajli et al. (2004) proposed a local bond stress–slip monotonic relationship to characterize
the behavior of steel bars when the splitting mode of bond failure occurs. Later Harajli (2009)
extended the work to account for cyclic bond degrading response, based on laboratory test data.
The model takes into account the level of confinement of the concrete and considers the effect of
critical bond parameters such as the diameter of steel bars, the ratio of concrete cover to bar
diameter, the concrete compressive strength, the type of confinement (steel ties, fiber-reinforced
concrete, or fiber-reinforced polymer jackets) and the area or content of confining reinforcement.
18

Zhao and Sritharan (2007) introduced a hysteretic model to capture strain penetration
effects. A reinforcing bar stress vs. slip response relationship was developed to be integrated into
fiber-based analysis of concrete structures using a zero-length section element. The model was
validated through comparison of results obtained when applying the model, with measured
global and local responses of two concrete columns and a bridge tee-joint system.

Chowdhury and Orakcal (2012) presented a modeling approach to apply in reinforced


concrete columns subjected to horizontal cyclic loading, where a bond stress vs. slip relationship
is implemented to springs in the formulation of a fiber-based flexural macro-model. This model
directly reflects the influence of local bond slip deformations due to the local coupling between
flexural deformations and slip deformations associated with either splitting or pullout bond
failures along the lap-splice region.

2.5 Soil-Structure Interaction of Shallow Foundations Structure Systems

Several authors addressed the effects of soil-structure interaction (SSI) on the structural
response, either analytically, numerically or through physical tests.

Regarding the analytical studies on shallow foundations, there is an evident interest within
the research and practicing engineering community on understanding how to deal with effect of
soil on structural response. The most common approaches used on modeling the soil-foundation
interaction are: (1) continuum finite element and boundary element approaches, (2) macro-
element formulations, and (3) Winkler-based approaches.

The continuum approach assumes a semi-infinite and isotropic material able to consider
multiple soil layers, however involves extensive computational effort (Drucker and Prager, 1952;
Fletcher and Hermann, 1971; Dutta and Roy, 2002).

Macro-element models consist of single element models, in which the soil-structure


interaction response is captured using plasticity-based formulations. Cremer et al. (2001), Gajan
(2006), Paolucci et al. (2008) and Figini et al. (2012) developed macro-element models for
shallow foundations. The main advantage with respect to other approaches is characterized by
the coupling between all the macro-element DOFs. On the other hand, the methodology to
19

develop the macro-element requires more complexity when compared with Winkler-based
approaches.

Winkler-based approaches are based on the utilization of 1-D spring elements that represent
the contact between the structural element and the soil (2-D or 3-D elements), representing
overall behavior of the soil-structure-interface. Winkler-based models have been commonly used
in the research community and several examples are provided next.

Chopra and Yim (1985) presented an analytical study evaluating the rocking response of a
single-degree-of-freedom (SDOF) system considering uplifting of the foundation, developing
simplified expressions to determine the base shear resistance of flexible structures allowed to
uplift. When applied to a multi-story building structure, the foundation flexibility and uplift had
considerable effect on the fundamental mode of vibration, although little effect on higher modes,
and can be critical for buildings with fundamental periods greater than two seconds.

Taylor et al. (1980) and Bartlett (1976) used Winkler-based models for capturing shallow
foundation rocking response where an analytical model for predicting moment-rotation behavior
of rigid footings using elastic-perfectly-plastic springs was coupled with Coulomb slider
elements. Elastic-plastic springs were considered to only have compression capacity, while
Coulomb slider elements captured the uplifting of the foundation.

Nakaki and Hart (1987) used Winkler springs with zero tension capacity and elastic
compressive resistance together with viscous dampers at the base of a shear wall structure. The
inelastic shear wall structure was modeled using a nonlinear stiffness degradation hysteretic
model. The fundamental period of the structure increases substantially and it was verified that
uplift of the foundation caused greater ductility demands on the structure.

Thomas et al. (2005) studied the nonlinear load-deformation characteristics during cyclic
and earthquake loading through five series of tests on a large centrifuge, including 40 models of
shear wall footings. Footing dimensions, depth of embedment, wall weight, initial static vertical
factor of safety, soil density, and soil type (dry sand and saturated clay) were systematically
varied. The moment capacity degradation was observed due to the deformed shape of the
footing–soil interface, along with the uplift associated with large rotations. Permanent
deformations beneath the footing continue to accumulate with the number of cycles of loading.
20

Allotey and Naggar (2008) adopted a Winkler-based modeling concept for capturing the
cyclic response of shallow foundations considering linear backbone curves developed in the
authors’ previous work (Allotey and El Naggar 2003). The model is able to predict the moment–
rotation and settlement response. However, the model is not able to capture the sliding response
adequately.

Rawchowdhury and Hutchinson (2008) performed experiments on shallow foundations in


order to calibrate the backbone curves for nonlinear springs of a Winkler-based modeling
framework that includes a distributed array of nonlinear inelastic springs, dashpots, and gap
elements. Different conditions were taken into account on the execution centrifuge experiments:
(1) square and strip footings; (2) bridge and building models; (3) static and dynamic loading, (4)
footings on sand and clay, (5) different static vertical factors of safety, and (6) different aspect
ratios. From all tests, was observed that the model can reasonably predict the footing response in
terms of moment, shear, settlement and rotational demands. The general hysteresis shape of the
moment- rotation, settlement-rotation and shear-sliding curves was also well captured.
Raychowdhury (2008) shows details on the model calibration and other important features of the
models.

Gajan et al. (2008; 2010) developed a contact interface model (CIM), which provides
nonlinear constitutive relations between cyclic loads and displacements at the footing-soil
interface of a shallow rigid foundation that is subjected to combined moment, shear, and axial
loading. This model assumes a rigid footing, and the interface model is able to couple the vertical
and horizontal responses. Both the CIM and BNWF models were applied to a shear wall building
resting on clayey foundation soils and to a shear wall and column building systems supported on
clean, dry, sand foundation soils tested in the centrifuge. Both the CIM and the BNWF models
captured the moment-rotation response consistent with the available experimental data. Their
shear-sliding behavior can deviate depending on the degree of foundation uplift.

Harden and Hutchinson (2009) adopted a BNWF simulation methodology for modeling
shallow foundation-structure systems, where seismically-induced rocking plays a predominant
role in the response. Numerical results indicated that a reasonable comparison between the
nonlinear Winkler-based approach and experimental response in terms of moment-rotation,
21

settlement-rotation, and shear-sliding displacement can be obtained, given an appropriate


selection of model and soil properties.

Gajan and Kutter (2009) performed several analyses using CIM and compared the results
with centrifuge model test results. The results indicated that the CIM is capable of capturing
essential features as such as load capacity, stiffness degradation, energy dissipation and
deformation of shallow foundation subjected to cyclic loading.

Raychowdhury (2011) applied a BNWF model on low-rise steel moment-resisting frame


(SMRF) buildings. The results obtained from fixed-base and elastic-base models were compared,
and it was observed that the force and displacement demands are reduced significantly when the
foundation nonlinearity is accounted for.

With respect to experimental tests performed to evaluate the SSI effects when using shallow
foundations, several studies are here addressed.

Deng and Kutter (2012) explored the behavior of rocking shallow foundation embedded in
dry sand. The results of slow (quasi-static) cyclic tests of rocking shear walls and dynamic
shaking tests of single-column rocking bridge models were presented. The behavior of rocking
foundation was shown to be sensitive to the geometric factor of safety with respect to bearing
failure, Lf/Lc, where Lf was the footing length, and the Lc was the critical soil-footing contact
length that would be required to support pure axial loading. In terms of settlements, they were
shown to be small if Lf/Lc was reasonably large.

Hakhamaneshi et al. (2012) reported a series of centrifuge model tests of shear walls and
SDOF structures with rocking foundations supported on clay of medium strength. Vertical
bearing failure, slow cyclic loading, and dynamic shaking tests were performed at three different
levels of acceleration. The measured moment capacity is found to match the theoretical values. A
settlement smaller than previous results for similar models tested on sand, was observed. The
greater settlements in sands are explained by the residual uplift associated with sand falling into
the gap under the footing as it rocked, occurred when facing very large rotations.

Mason et al. (2013) and Trombetta et al. (2013) studied the seismic soil–foundation–
structure interaction (SFSI), and structure–soil–structure interaction (SSSI) of frame structures.
22

The SSSI was considered for the case when the shallow foundations of a building are adjacent to
a deep basement of other building. Results showed that the presence of the deep basement affects
the moment–rotation behavior of the adjacent shallow footings, stiffening the response in the
direction of loading towards the basement, limiting the permanent displacements of the footing,
which in turn limits physical damage to the superstructure. These studies suggested that SSSI
should also be considered in the design of closely clustered structures.

2.6 Damage Indices

Structural damage can be quantified using damage indices that are functions of deformation,
ductility, stiffness degradation, and energy dissipated, among others. Damage indices available
in the literature can be classified as (1) deformation-based, (2) energy-based, or a combination of
both, i.e., (3) deformation- and energy-based indices.

For the deformation-based damage indices, Veletsos and Newmark (1960) first proposed a
damage index based on the peak inelastic deformation. Roufaiel and Meyer (1987) proposed
global damage measures based on peak roof displacement and roof displacement at failure of
Reinforced Concrete moment frames. Banon et al. (1981) and Lybas and Sozen (1977)
developed damage indices in which the concepts of cyclic deterioration were incorporated in the
analysis of structural systems consisting of reinforced concrete moment frames and reinforced
concrete coupled walls, respectively. Stephens and Yao (1987) developed a cumulative damage
index based on displacement ductility, after testing two concrete structures and studying the
correlation between the damage observed in both. Jeong and Iwan (1988) proposed a classical
low-cycle fatigue formulation calibrating their model with observed failures of reinforced
concrete columns. Chung et al. (1987) extended the concept of damage and structural failures
and developed a damage index formulation that also accounts for fatigue.

In terms of energy based damage indices, Gosain et al. (1977) developed a cumulative
energy ratio for reinforced concrete members. Darwin and Nmai (1986) applied a factor to the
damage index developed by Gosain et al. in order to account for the reinforcing steel layout of
reinforced concrete beams. Kratzig et al. (1989) proposed a more complex energy formulation
based on the definition of primary and follower half-cycles for reinforced concrete members. To
quantify the structural damage, McCabe and Hall (1989) considers the energy dissipated as the
23

main component in the evaluation of two damage measures, one based on the concept of
equivalent hysteretic cycles of deformation and the second on concepts derived from low‐cycle
fatigue theory. Fajfar (1992) and Powel and Allahadabi (1988) also used the number of cycles, as
well as the amplitude or history of the inelastic cyclic response, to define two other damage
indices.

Regarding deformation- and energy-based combined damage indices, Bracci et al. (1989)
proposed a combined approach that defined the potential damage as the total area between the
monotonic load-deformation and the fatigue failure envelope. DiPasquale and Cakmak (1990)
presented yet another procedure to evaluate the damage in the structure that is related to the
overall stiffness loss in the structure due to inelastic response through tracking the elongation of
the fundamental period of vibration. Park and Ang (1985) proposed an index based on the
combination between maximum deformations in the structure and the effect of the repetition of
load cycles through a term that considers the total energy dissipated and the energy dissipated
during the loading cycles. Reinhorn and Valles (1996) extended the Park and Ang damage index
(DIP&A) to account for a damage measure that was calibrated taking into consideration an
empirically calibrated fatigue-based damage model.

In terms of defining damage states based on the values of the damage indices, Park et al.
(Park et al. 1987) calibrated the damage index proposed by Park and Ang (1985) with damage
observed in nine reinforced concrete buildings damaged by the 1971 San Fernando and 1978
Miyagiken-Oki earthquakes. The proposed breakdown for classifying damage states based on
this work is shown. The same breakdown was later adopted for the Reinhorn and Valles damage
index (DIR&V). Stone and Taylor (1993) proposed new limits for the damage states based on the
Park and Ang damage index to use in circular bridge columns, and Ang et al. (1993) proposed
further adjustments to use in bridge structures.

2.7 Unresolved questions on the topic

Several questions arise when the ground motion duration is linked with other topics
reviewed in this chapter such as:
24

 Seismic assessment of irregular buildings, where the torsional behavior plays an


important role on the seismic assessment of the building;
 Seismic assessment of bridges taking into account: (1) a retrofitting solution applied on
their columns and the impact of that solution on the role played by ground motion
duration effects, and (2) irregularity on their geometry, characterized by having columns
with different heights along the bridge, while varying the columns position.
 Effects of SSI on the structural response, understanding how the role played by ground
motion duration is affected when considering SSI.

This work gives a significant contribution to the objective of filling these gaps in knowledge
with the studies presented in chapters 3, 4 and 5.
25

2.8 Bibliography

AASHTO, L. (2012). “Bridge Design Specifications 2012 Washington.” DC, USA American
Association of State Highway and Transportation Officials.

Abedi, K., Afshin, H., and Shirazi, M. R. N. (2010). “Numerical study on the seismic retrofitting
of reinforced concrete columns using rectified steel jackets.” Asian Journal of Civil
Engineering (Building and Housing), 11(2), 219–240.

Allotey, N., and El Naggar, M. H. (2003). “Analytical moment-rotation curves for rigid
foundations based on a Winkler model.” Soil Dynamics and Earthquake Engineering, 23(5),
367–381.

Allotey, N., and El Naggar, M. H. (2008). “An investigation into the Winkler modeling of the
cyclic response of rigid footings.” Soil Dynamics and Earthquake Engineering, 28(1), 44–
57.

Amneus, D. (2015). “Methods for Strengthening Flexural Steel Details in Reinforced Concrete
Bridge Girders Using a Near-surface Mounted Retrofitting Technique.” Oregon State
University.

Ang, A. H.-S., Kim, W. J., and Kim, S. B. (1993). “Damage Estimation of Existing Bridge
Structures.” Structural Engineering in Naturakl Hazards Mitigation:ASCE Structures
Congress, Irvine, CA, 1137–1142.

Banon, H., Irvine, H. M., and Biggs, J. M. (1981). “Seismic damage in reinforced concrete
frames.” Journal of the Structural Division, ASCE, 107(9), 1713–1729.

Barbosa, A. R., Ribeiro, F. L. A., and Neves, L. C. (2017). “Influence of earthquake ground-
motion duration on damage estimation: application to steel moment resisting frames.”
Earthquake Engineering and Structural Dynamics, 46(1), 27–49.

Barker, L. M. (2015). “Flexural Anchorage Performance and Strengthening on Negative Moment


Regions Using Near-surface Mounted Retrofitting in Reinforced Concrete Bridge Girders.”
Oregon State University.

Bartlett, P. E. (1976). Foundation rocking on clay soil. Report (University of Auckland, School of
Engineering). Department of Civil Engineering, University of Auckland.

Bazaez, R., and Dusicka, P. (2016a). “Cyclic Loading for RC Bridge Columns Considering
Subduction Megathrust Earthquakes.” Journal of Bridge Engineering, American Society of
Civil Engineers, 21(5), 4016009.
26

Bazaez, R., and Dusicka, P. (2016b). “Cyclic behavior of reinforced concrete bridge bent
retrofitted with buckling restrained braces.” Engineering Structures, Elsevier, 119, 34–48.

Bojorquez, E., Iervolino, I., Manfredi, G., and Cosenza, E. (2006). “Influence of ground motion
duration on degrading SDOF systems.” 1st European Conference on Earthquake
Engineering and Seismology, (September), 3–8.

Bommer, J. J., Magenes, G., Hancock, J., and Penazzo, P. (2004). “The Influence of Strong-
Motion Duration on the Seismic Response of Masonry Structures.” Bulletin of Earthquake
Engineering, 2(1), 1–26.

Bozorgnia, Y., and Tso, W. K. (1984). “Inelastic earthquake response of asymmetric structures.”
ASCE Journal of Structural Engineering, 112(2), 383–400.

Bracci, J. M., Reinhorn, A. M., Mander, J. B., and Kunnath, S. K. (1989). Deterministic model
for seismic damage evaluation of RC Structures. Technical Report NCEER-89-0033.

Cairns, J., and Arthur, P. D. (1979). “Strength of lapped splices in reinforced concrete columns.”
Journal Proceedings, 76(2), 277–296.

Chai, Y. H. (2005). “Incorporating low-cycle fatigue model into duration-dependent inelastic


design spectra.” Earthquake Engineering and Structural Dynamics, 34(1), 83–96.

Chakroborty, S., and Roy, R. (2013). “Role of Ground Motion Characteristics on Inelastic
Seismic Response of Irregular Structures.” 22(1), 1–16.

Chandramohan, R., Baker, J. W., and Deierlein, G. G. (2016). “Quantifying the Influence of
Ground Motion Duration on Structural Collapse Capacity Using Spectrally Equivalent
Records.” Earthquake Spectra, 32(2), 927–950.

Chang, S.-Y., Li, Y.-F., and Loh, C.-H. (2004). “Experimental study of seismic behaviors of as-
built and carbon fiber reinforced plastics repaired reinforced concrete bridge columns.”
Journal of Bridge Engineering, American Society of Civil Engineers, 9(4), 391–402.

Chopra, A. K., and Yim, S. C. ‐S. (1985). “Simplified Earthquake Analysis of Structures with
Foundation Uplift.” Journal of Structural Engineering, 111(4), 906–930.

Chowdhury, S. R., and Orakcal, K. (2012). “An analytical model for reinforced concrete
columns with lap splices.” Engineering Structures, Elsevier, 43, 180–193.

Chung, Y. S., Meyer, C., and Shinozuka, M. (1987). Seismic damage assessment of RC
members. Technical Report NCEER-87-0022.

Cornell, A. C. (1997). “Does duration really matter?” Proceedings of the FHWA/NCEER


Workshop on the National Representation of Seismic Ground Motion for New and Existing
27

Highway Facilities, Burlingame, Calif., organized by NCEER (project 106-F-5.4.1) and


ATC (project ATC-18-1), 125–133.

Cremer, C., Pecker, A., and Davenne, L. (2001). “Cyclic macro-element for soil-structure
interaction: Material and geometrical non-linearities.” International Journal for Numerical
and Analytical Methods in Geomechanics, 25(13), 1257–1284.

Darwin, D., and Nmai, C. (1986). “Energy dissipation in RC beams under cyclic load.” ASCE
Journal of Structural Engineering, 112(8), 1829–1846.

De-la-Colina, J. (2003). “Assessment of design recommendations for torsionally unbalanced


multistory buildings.” Earthquake Spectra, 19(1), 47–66.

Deng, L., and Kutter, B. L. (2012). “Characterization of rocking shallow foundations using
centrifuge model tests.” Earthquake Engineering and Structural Dynamics, 41(5), 1043–
1060.

DiPasquale, E., and Cakmak, A. S. (1990). “Relation between Global Damage Indices and Local
Stiffness Degradation.” ASCE Journal of Structural Engineering.

Drucker, D. C., and Prager, W. (1952). “Soil mechanics and plastic analysis or limit design.”
Quarterly of applied mathematics, 10(2), 157–165.

Dusicka, P., and Lopez, A. (2016). Impact of Cascadia Subduction Zone Earthquake on the
Seismic Evaluation Criteria of Bridges. Portland.

Dutta, A., and Mander, J. B. (2001). “Energy Based Methodology for Ductile Design pf
Concrete Columns.” ASCE Journal of Structural Engineering, 127(12), 1374–1381.

Dutta, S. C., and Das, P. K. (2002). “Inelastic seismic response of code-designed reinforced
concrete asymmetric buildings with strength degradation.” Engineering Structures, 24(10),
1295–1314.

Dutta, S. C., and Roy, R. (2002). “A critical review on idealization and modeling for interaction
among soil--foundation--structure system.” Computers & structures, Elsevier, 80(20),
1579–1594.

ElGawady, M., Endeshaw, M., McLean, D., and Sack, R. (2010). “Retrofitting of Rectangular
Columns with Deficient Lap Splices.” Journal of Composites for Construction, 14(1), 22–
35.

EN 1998-1. (2004). “Eurocode 8: Design of structures for earthquake resistance—Part 1: General


rules, seismic actions and rules for buildings.” European Committee for Normalization,
Brussels.
28

Endeshaw, M. a., ElGawady, M., Sack, R. L., and McLean, D. I. (2008). Retrofit of Rectangular
Bridge Columns using CFRP Wrapping.

Faggella, M., Barbosa, A. R., Conte, J. P., Spacone, E., and Restrepo, J. I. (2013). “Probabilistic
seismic response analysis of a 3-D reinforced concrete building.” Structural Safety, 44, 11–
27.

Fajfar, P. (1992). “Equivalent ductility factors, taking into account low-cycle fatigue.”
Earthquake Engineering and Structural Dynamics, 21(10), 837–848.

Feng, M. Q., Bhatia, H., Baird, K., and Elsanadedy, H. (1999). Structural Qualification Testing
of Composite-jacketed Circular and Rectangular Bridge Columns. Irvine, CA.

Figini, R., Paolucci, R., and Chatzigogos, C. T. (2012). “A macro-element model for non-linear
soil--shallow foundation--structure interaction under seismic loads: theoretical development
and experimental validation on large scale tests.” Earthquake Engineering & Structural
Dynamics, Wiley Online Library, 41(3), 475–493.

Fletcher, D. Q., and Hermann, L. R. (1971). “Elastic foundation representation of continuum.”


Journal of the Engineering Mechanics Division, ASCE, 97(1), 95–107.

Foschaar, J. C., Baker, J. W., and Deierlein, G. G. (2012). “Preliminary Assessment of Ground
Motion Duration Effects on Structural Collapse.” Proceedings of 15th World Conference on
Earthquake Engineering, Lisbon, Portugal.

Gajan, S. (2006). “Physical and numerical modeling of nonlinear cyclic load-deformation


behavior of shallow foundations supporting rocking shear walls.”

Gajan, S., and Kutter, B. L. (2008). “Capacity, Settlement, and Energy Dissipation of Shallow
Footings Subjected to Rocking.” Journal of Geotechnical and Geoenvironmental
Engineering, 134(8), 1129–1141.

Gajan, S., and Kutter, B. L. (2009). “Contact Interface Model for Shallow Foundations Subjected
to Combined Cyclic Loading.” Journal of Geotechnical and Geoenvironmental
Engineering, 135(3), 407–419.

Gajan, S., Raychowdhury, P., Hutchinson, T. C., Kutter, B. L., and Stewart, J. P. (2010).
“Application and validation of practical tools for nonlinear soil-foundation interaction
analysis.” Earthquake Spectra, 26(1), 111–129.

Gidaris, I., Padgett, J. E., Barbosa, A. R., Chen, S., Cox, D., Webb, B., and Cerato, A. (2016).
“Multiple-Hazard Fragility and Restoration Models of Highway Bridges for Regional Risk
and Resilience Assessment in the United States: State-of-the-Art Review.” Journal of
Structural Engineering, 04016188.
29

Girard, C., and Bastien, J. (2002). “Finite-Element Bond-Slip Model for Concrete Columns
under Cyclic Loads.” Journal of Structural Engineering, 128(12), 1502–1510.

Gosain, N. K., Moore, W. P., and Jirsa, J. O. (1977). “Shear Requirements for Load Reversals on
RC Members.” Journal of the Structural Division, ASCE, 103(7), 1461–1476.

Hakhamaneshi, M., Kutter, B. L., Deng, L., Hutchinson, T. C., and Liu, W. (2012). “New
findings from centrifuge modeling of rocking shallow foundations in clayey ground.”
GeoCongress 2012: State of the Art and Practice in Geotechnical Engineering, 195–204.

Hancock, J., and Bommer, J. J. (2006). “A state-of-knowledge review of the influence of strong-
motion duration on structural damage.” Earthquake Spectra, 22(3), 827–845.

Harajli, M. H. (2009). “Bond stress--slip model for steel bars in unconfined or steel, FRC, or
FRP confined concrete under cyclic loading.” Journal of Structural Engineering, American
Society of Civil Engineers, 135(5), 509–518.

Harajli, M. H., Hamad, B. S., and Rteil, A. A. (2004). “Effect of Confinement of Bond Strength
between Steel.” ACI Structural Journal, 101(5).

Harden, C. W., and Hutchinson, T. C. (2009). “Beam-on-nonlinear-winkler-foundation modeling


of shallow, rocking-dominated footinǵs.” Earthquake Spectra, 25(2), 277–300.

Iacobucci, R. D., Sheikh, S. A., and Bayrak, O. (2003). “Retrofit of square concrete columns
with carbon fiber-reinforced polymer for seismic resistance.” ACI Structural Journal,
American Concrete Institute, 100(6), 785–794.

Iervolino, I., Manfredi, G., and Cosenza, E. (2006). “Ground motion duration effects on
nonlinear seismic response.” Earthquake Engineering and Structural Dynamics, 35(1), 21–
38.

International Code Council. (1997). “Uniform building code (IBC).” International Conference of
Building Officials.

Jeong, G. D., and Iwan, W. D. (1988). “The effect of earthquake duration on the damage of
structures.” Earthquake Engineering and Structural Dynamics, 16(8), 1201–1211.

Kappos, A. J., Manolis, G. D., and Moschonas, I. F. (2002). “Seismic assessment and design of
R/C bridges with irregular congiguration, including SSI effects.” Engineering Structures,
24(10), 1337–1348.

Katsumata, H., Kimura, K., and Kobatake, Y. (1998). “Seismic Retrofitting Technique Using
Carbon Fibers For Reinforced Concrete Buildings.” Strain, 2(4), 0.
30

Kayen, R. E., and Mitchell, J. K. (1997). “Assessment of liquefaction potential during


earthquakes by Arias intensity.” Journal of Geotechnical and Geoenvironmental
Engineering, American Society of Civil Engineers, 123(12), 1162–1174.

Kim, S.-H., and Shinozuka, M. (2004). “Development of fragility curves of bridges retrofitted by
column jacketing.” Probabilistic Engineering Mechanics, Elsevier, 19(1), 105–112.

Kramer, S. L., and Mitchell, R. A. (2006). “Ground motion intensity measures for liquefaction
hazard evaluation.” Earthquake Spectra, 22(2), 413–438.

Krätzig, W. B., Meyer, I. F., and Meskouris, K. (1989). “Damage Evolution in Reinforced
Concrete Members Under Cyclic Loading.” 5th International Conference on Structural
Safety and reliability (ICOSSAR 89) - Vol II, San Francisco CA, 795–802.

Lostra, M. (2016). “Seismic Performance of Square Reinforced Concrete Columns Retrofitted


with Titanium Alloy Bars.” Oregon State University.

Lostra, M., Higgins, C., and Barbosa, A. R. (2017). “Seismic retrofit of reinforced concrete
bridge columns using titanium-alloy bars.” Portuguese Journal of Structural Engineering,
3, 75–82.

Lukose, K., Gergely, P., and White, R. N. (1982). “Behavior of reinforced concrete lapped
splices for inelastic cyclic loading.” Journal Proceedings, 355–365.

Lybas, J. ., and Sozen, M. (1977). Effect of beam strength ratio on dynamic behaviour of
reinforced concrete coupled walls. Report SRS No. 444.

Lynn, a C., Moehle, J. P., Mahin, S. a, and Holmes, W. T. (1996). “Seismic evaluation of
existing reinforced concrete building columns.” Earthquake Spectra.

Mackie, K. R., and Stojadinović, B. (2005). Fragility basis for California highway overpass
bridge seismic decision making. Pacific Earthquake Engineering Research Center, College
of Engineering, University of California, Berkeley, Berkeley, California.

Mackie, K. R., and Stojadinović, B. (2006). “Post-earthquake functionality of highway overpass


bridges.” Earthquake engineering and structural dynamics, Wiley Online Library, 35(1),
77–93.

Mander, J. B., Priestley, M. J. N., and Park, R. (1988). “Theoretical stress-strain model for
confined concrete.” Journal of structural engineering, American Society of Civil Engineers,
114(8), 1804–1826.

Manfredi, G., Polese, M., and Cosenza, E. (2003). “Cumulative demand of the earthquake
ground motions in the near source.” Earthquake Engineering and Structural Dynamics,
32(12), 1853–1865.
31

Mason, H. B., Trombetta, N. W., Chen, Z., Bray, J. D., Hutchinson, T. C., and Kutter, B. L.
(2013). “Seismic soil--foundation--structure interaction observed in geotechnical centrifuge
experiments.” Soil Dynamics and Earthquake Engineering, Elsevier, 48, 162–174.

Mccabe, S. L., and Hall, W. J. (1989). “Assessment od Seismic Structural Damage.” ASCE
Journal of Structural Engineering, 115(9), 2166–2183.

McLean, D., and Bernards, L. (1992). Seismic Retrofitting of Rectangular Bridges Column for
Shear. Pullman, Washington.

Melek, M., and Wallace, J. W. (2004). “Cyclic behavior of columns with short lap splices.” ACI
Structural Journal, 101(6), 802–811.

Moehle, J. P., and bavanagh, T. (1985). “Confinement effectiveness of crossties in RC.” Journal
of Structural Engineering, American Society of Civil Engineers, 111(10), 2105–2120.

Mohammadi-Tamanani, M., Gian, Y., and Ayoub, A. (2014). “Design of bridges with unequal
pier heights.” Structures Congress 2014, Boston, Massachusetts, 677–686.

Nakaki, D. K., and Hart, G. C. (1987). “Uplifting response of structures subjected to earthquake
motions.” EKEH.

Nassar, A., and Krawinkler, H. (1991). Seismic Demands for {SDOF} and {MDOF} Systems.
California.

Nielson, B. G., and DesRoches, R. (2007a). “Seismic fragility methodology for highway bridges
using a component level approach.” Earthquake Engineering & Structural Dynamics, 36,
823–839675.

Nielson, B. G., and DesRoches, R. (2007b). “Analytical seismic fragility curves for typical
bridges in the central and southeastern United States.” Earthquake Spectra, 23(3), 615–633.

Oyarzo-vera, C., and Chouw, N. (2008). “Effect of earthquake duration and sequences of ground
motions on structural responses.” Proceedings of the 10th International Symposium on
Structural Engineering of Young Experts, Changsha, Scottland.

Padgett, J. E., and DesRoches, R. (2009). “Retrofitted bridge fragility analysis for typical classes
of multispan bridges.” Earthquake Spectra, 25(1), 117–141.

Paolucci, R., Shirato, M., and Yilmaz, M. T. (2008). “Seismic behaviour of shallow foundations:
Shaking table experiments vs numerical modelling.” Earthquake Engineering & Structural
Dynamics, Wiley Online Library, 37(4), 577–595.

Park, Y., and Ang, A. (1985). “Mechanistic Seismic Damage Model for Reinforced Concrete.”
ASCE Journal of Structural Engineering, 111(4), 722–739.
32

Park, Y. J., Ang, A. H. S., and Wen, Y. K. (1987). “Damage-Limiting Aseismic Design of
Buildings.” Earthquake Spectra.

Paulay, T. (1982). “Lapped splices in earthquake-resisting columns.” Journal Proceedings, 458–


469.

Paulay, T. (1997). “A Review of Code Provisions for Torsional Seismic Effects in Buildings.”
Bulletin of the New Zeland National Society for Earthquake Engineering, 30(3).

Pauley, T., and Priestley, M. J. N. (1992). Seismic Design of Reinforced Concrete and Masonry
Buildings. Wiley.

Pinho, R., Monteiro, R., Casarotti, C., and Delgado, R. (2009). “Assessment of continuous span
bridges through nonlinear static procedures.” Earthquake Spectra, 25(1), 143–159.

Powell, G. H., and Allahabadi, R. (1988). “Seismic damage prediction by deterministic methods:
Concepts and procedures.” Earthquake Engineering and Structural Dynamics, 16(5), 719–
734.

Priestley, M. J. N., Seible, F., and Calvi, G. M. (1994). Seismic assessment of existing bridges.
Wiley Online Library.

Raghunandan, M., and Liel, A. B. (2013). “Effect of ground motion duration on earthquake-
induced structural collapse.” Structural Safety, Elsevier, 41, 119–133.

Raghunandan, M., Liel, A. B., and Luco, N. (2015). “Collapse Risk of Buildings in the Pacific
Northwest Region due to Subduction Earthquakes.” Earthquake Spectra, 31(4), 2087–2115.

Raychowdhury, P. (2008). “Nonlinear winkler-based shallow foundation model for performance


assessment of seismically loaded structures.” University of California, San Diego.

Raychowdhury, P. (2011). “Seismic response of low-rise steel moment-resisting frame (SMRF)


buildings incorporating nonlinear soil-structure interaction (SSI).” Engineering Structures,
Elsevier Ltd, 33(3), 958–967.

Raychowdhury, P., and Hutchinson, T. (2008). “Nonlinear material models for Winkler-based
shallow foundation response evaluation.” GeoCongress 2008: Characterization,
Monitoring, and Modeling of GeoSystems, New Orleans, Louisiana, 9–12.

Roufaiel, M. S. L., and Meyer, C. (1987). “Analytical Modeling of Hysteretic Behavior of R/C
Frames.” ASCE Journal of Structural Engineering, 113(3), 429–444.

Ruiz-García, J. (2010). “On the influence of strong-ground motion duration on residual


displacement demands.” Earthquake and Structures, 1(4), 327–344.
33

Schlick, B. M., and Breña, S. F. (2004). “Seismic rehabilitation of reinforced concrete bridge
columns in moderate earthquake regions using FRP composites.” Proc. of the 13th World
Conference on Earthquake Engineering,(CD-Rom, paper n. 508), Vancouver, BC, Canada,
August, 1–6.

Seible, F., Priestley, M. J. N., Hegemier, G. a., and Innamorato, D. (1997). “Seismic Retrofit of
RC Columns with Continuous Carbon Fiber Jackets.” Journal of Composites for
Construction, 1(2), 52–62.

Sheikh, S. A., and Uzumeri, S. M. (1980). “Strength and ductility of tied concrete columns.”
Journal of the structural division, 106(ASCE 15388 Proceeding).

Shome, N., Cornell, C. A., and Bazurro, P. (1998). “Earthquakes, records, and nonlinear
responses.” Earthquake Spectra, 14(3), 469–500.

Sideras, S. S., and Kramer, S. L. (2012). “Potential implications of long duration ground motions
on the response of liquefiable soil deposits.” 9th International Conference on Urban
Earthquake Engineering/4th Asia Conference on Earthquake Engineering, Tokyo, Japan.

Sommer, A., and Bachmann, H. (2005). “Seismic behavior of asymmetric RC wall buildings:
Principles and new deformation-based design method.” Earthquake Engineering and
Structural Dynamics, 34(2), 101–124.

Song, R., Li, Y., and van de Lindt, J. W. (2014). “Impact of earthquake ground motion
characteristics on collapse risk of post-mainshock buildings considering aftershocks.”
Engineering Structures, Elsevier Ltd, 81, 349–361.

Stathopoulos, K. G., and Anagnostopoulos, S. a. (2005). “Inelastic torsion of multistorey


buildings under earthquake excitations.” Earthquake Engineering and Structural Dynamics,
34(12), 1449–1465.

Stephens, J. E., and Yao, J. T. P. (1987). “Damage assessment using response measurements.”
ASCE Journal of Structural Engineering, 113(4), 787–801.

Stone, W. C., and Taylor, A. W. (1993). Seismic performance of circular bridge columns
designed in accordance with AASHTO/CALTRANS standards. National Institute of
Standards and Technology.

Taylor, P. W., Bartlett, P. E., and Wiessing, P. R. (1980). Foundation rocking under earthquake
loading. University of Auckland, Department of Civil Engineering.

Tena-colunga, A. (2004). “Evaluation of the seismic response of slender , setback RC moment-


resisting frame buildings designed according to the seismic guidelines of a modern seismic
code.” Proceedings of the 13th World Conference on Earthquake Engineering, Vancouver,
Canada.
34

Thomas, J., Gajan, S., and Kutter, B. (2005). “Soil-Foundation-Structure Interaction: Shallow
Foundations. Centrifuge Data Report for the SSG04 Test Series.” … for Geotechnical
Modeling Data Report …, (June 2003).

Tremblay, R., and Atkinson, G. M. (2001). “Comparative Study of the Inelastic Seismic Demand
of Eastern and Western Canadian Sites.” Eartquake Spectra, 17(2), 333–358.

Trombetta, N. W., Mason, H. B., Chen, Z., Hutchinson, T. C., Bray, J. D., and Kutter, B. L.
(2013). “Nonlinear dynamic foundation and frame structure response observed in
geotechnical centrifuge experiments.” Soil Dynamics and Earthquake Engineering,
Elsevier, 50, 117–133.

Valles, R. E., Reinhorn, A. M., Kunnath, S. K., Li, C., and Madan, A. (1996). IDARC 2D
Version 4.0: A Program for the Inelastic Damage Analysis of Buildings. Technical Report
NCEER-96-0010. Technical Report NCEER-96-0010.

Veletsos, A., and Newmark, N. M. (1960). “Effect of inelastic behavior on the response of
simple systems to earthquake motions.” Proceedings of the 2nd World Conference on
Earthquake Engineering.

Watson, S, Zahn, F.A., Park, R. (1994). “achieve specified ductility levels in the potential
plastic-hinge regions of columns . The derived equations more accurately reflect the
influence of the significant variables than current code equations . Also , the enhancement
of the flexural strength o.” 120(6), 1798–1824.

Zhang, Y., Conte, J. P., Yang, Z., Elgamal, A., Bielak, J., and Acero, G. (2008). “Two-
dimensional nonlinear earthquake response analysis of a bridge-foundation-ground system.”
Earthquake Spectra, 24(2), 343–386.

Zhao, J., and Sritharan, S. (2007). “Modeling of strain penetration effects in fiber-based analysis
of reinforced concrete structures.” ACI structural journal, American Concrete Institute,
104(2), 133.
35

3 INFLUENCE OF GROUND MOTION DURATION ON DAMAGE INDEX-BASED


FRAGILITY ASSESSMENT OF A PLAN-ASYMMETRIC NON-DUCTILE
REINFORCED CONCRETE BUILDING

Andre Belejo, Andre Ramos Barbosa, and Rita Bento

EQUATION CHAPTER 3 SECTION 1

Accepted for publication in


Engineering Structures
ELSEVIER
36

3.1 Abstract

The role of ground motion duration on the seismic performance of building structures
remains unclear. This paper presents results on the effects of ground motion duration on the
seismic behavior of a 3-D plan-asymmetric reinforced concrete building. A three-story
reinforced concrete building tested in Europe is used as a case study. A nonlinear model of the
building with fiber-section distributed inelasticity displacement-based beam-column elements is
subjected to two ground motion sets: (1) long-duration bi-directional earthquake ground motions,
and (2) short-duration bi-directional earthquake ground motions. A new ground motion selection
procedure is proposed to isolate the effect of duration on the structural response and on damage
assessment. For comparative studies, long- and short-duration ground motions are selected so
that they have similar response spectra. The damage assessment is performed analyzing global
and local responses. Global responses of the structure are evaluated in terms of peak roof drift
ratios, peak interstory drift ratios (IDRs), IDR envelopes, normalized roof displacements relative
to motion of the center of gravity of the roof, and dissipated energy. Local responses assessed
include moment-curvature or moment-rotation response in beams and columns as well as the
hysteretic energy dissipated at selected sections of the structure. Two damage indices available in
the literature are used to characterize the level of damage that the structure experiences, through
damage index-based fragility curves that are developed as part of the work. For the short-
duration motions, fragility curves developed independently from interstory responses and from
the damage indices are similar, thus providing confidence into the use of the damages indices for
fragility curve development. In addition, results indicate that for the vintage plan-asymmetric
building analyzed, the ground motion duration plays a role in the damage predicted using the two
damage indices.

3.2 Introduction

Recent great earthquakes such as the Chile, Maule 2010 (Mw = 8.8) or Tohoku, Japan 2011
(Mw = 9.0) caused significant structural damage in buildings and bridges. The ground motions
recorded in seismic stations during these events are characterized by having significantly longer
duration than records typically recorded from major earthquakes (Mw = 7.0 to 7.9) in other
37

locations in the world. Other events such as the 1700 Cascadia Subduction Zone Great
Earthquake (~M9.0) in the United States, and the great Lisbon earthquake of November 1st, 1755
(~M8.7) are also expected to have produced similar long ground motion durations and in both
regions there is a significant probability that a great earthquake may strike in the next 50 years
(Cohee et al. 1991; Chester 2001; Raghunandan et al. 2015).

Several studies over the past decades have studied the effect of ground motion duration on
structural response, often with inconclusive results (Nassar and Krawinkler 1991; Shome et al.
1998; Tremblay and Atkinson 2001; Chai 2005). Results in other studies (Cornell 1997; Bommer
et al. 2004; Bojorquez et al. 2006; Hancock and Bommer 2006; Iervolino et al. 2006) indicated
that there is no significant correlation between measures of significant duration and peak IDRs in
building structures. However, the lack of capability of structural models in capture the effect of
degrading energy capacity and in-cycle strength degradation, and the lack of long-duration
ground motion records, hindered the possibility of addressing the effect of ground motion
duration in previous research.

Barbosa et al. (2017) and Chandramohan et al. (2016) used nonlinear finite element models
that captured strength and stiffness degradation, and presented results showing that for large
intensities, long-duration ground motions can induce higher deformations. Other studies that
focused on cumulative response measures such as (Dutta and Mander 2001; Manfredi et al.
2003; Oyarzo-vera and Chouw 2008) defend that the amount of accumulated damage depends on
the duration; Ruiz-Garcia (2010) observed larger residual deformations induced by long-duration
ground motions; Song et al. (2014), Foschaar et al. (2012) and Raghunandan and Liel (2013)
highlighted the importance of the ground motion duration, in which a higher probability of
collapse is observed after analyzing buildings subjected to long-duration ground motions when
compared with to short-duration ground motions. Raghunandan et al. (2015) performed a
collapse assessment of structures subjected to crustal and subduction zone ground motions
(which typically exhibited longer durations); results indicated that there was a greater risk of
structural collapse for subduction zone ground motions. Results in Chandramohan et al. (2016)
also indicate that higher probability of structural collapse are expected when structures are
subjected to long-duration motions. Barbosa et al. (2017) quantified the influence of ground
motion duration on the damage of a 3-, 9-, and 20-story steel building using two damage indices;
38

results shown indicated that the increase in energy dissipation demands observed in the
structures subjected to long-duration motions increased significantly the expected levels of
structural damage. In the assessment of robustness, mainshock-aftershock sequences and the
impact of the structure being subjected to long-duration shaking was also observed (Ribeiro et al.
2014).

It is worth noting that most studies discussed in the previous paragraph were performed
using one-dimensional (1-D) or two-dimensional (2-D) nonlinear structural models. To the best
to the authors’ knowledge no studies have addressed the assessment of ground motion duration
using 3-D plan-asymmetric buildings. Nonetheless, despite the fact that plan asymmetry (also
known as torsional irregularity) of strength and stiffness in building is known to produce larger
demands on some elements of the building, which in turn lead to often premature and unexpected
failures, in the existing body of literature on the effects of ground motion duration, a limited
number of studies were identified in which the effects of ground motion duration have been
studied using the three-dimensional (3-D) structural models of irregular structures (Chakroborty
and Roy 2013). For plan-asymmetric buildings, however, it is well known that localized damage
is a characteristic of this type of buildings under earthquakes. Bozorgnia and Tso (1984)
concluded that the effect of asymmetry on the element ductility demand and on edge
displacements plays an important role for stiff systems, reducing the elements yield strength.
Paulay (1997) described how the asymmetry of mass, stiffness and strength distributions that
cause torsional behavior in irregular buildings, makes it impossible to reproduce the seismic
effects using 2-D models, indicating that 3-D building models should always be used.
Stathopoulos and Anagnostopoulos (2005) showed that response measures as ductility factors
and damage indices in the flexible side of plan asymmetric buildings are up to two times greater
than those at the stiff side leading to premature member failures. Other authors (Dutta and Das
2002; De-la-Colina 2003; Tena-colunga 2004; Sommer and Bachmann 2005), performed seismic
assessments of 3-D structural models of code-designed plan irregular structures and concluded
that demands in irregular structures far exceeded the demands estimated using codified
equations. Faggella et al. (2013) illustrated the effects of scaling of components in two
orthogonal directions and the impact on the structural demands of an irregular RC moment frame
building structure. With respect to ground motion duration on the response of plan-asymmetric
39

structures, results in Roy and Chakroborty (2013) indicate that seismic vulnerability of plan-
asymmetric structures is sensitive to the characteristics of ground motions.

The objective of this study is to advance the understanding of the effects of ground motion
duration on the response of plan-asymmetric RC building structures when subjected to long-
duration and short-duration earthquake ground motions. A plan asymmetric three-story, as-built,
full-scale RC building structure tested in Europe is used as a case study. This building represents
a real 3-story building located in Southern Europe that was designed only for gravity loads based
on the design and construction practice applied in the early 1970s in the region, thus not being a
seismically design structure. A three-dimensional nonlinear finite element model of the three-
story structure is developed in OpenSees (McKenna et al. 2010). The finite element model
developed is validated using data available in the literature (Negro et al. 2004; Fardis and Negro
2005). A database of 64 ground motions, corresponding to 32 long- and 32 short-duration
spectrally equivalent bi-directional ground motion records, is used in this work to assess the
effects of ground motion duration on the structural behavior. An earthquake is classified as a
long-duration earthquake when at least one of its orthogonal components has a significant
duration, D5-75, greater than 25 seconds, which follows the definition of long-duration proposed
by Chandramohan et al. (2016) and Barbosa et al. (2017). The significant duration D5-75 of a
ground motion is defined as the time interval between accumulated Arias Intensity (Ia) of 5% to
75%, as represented in Figure 3.1.
40

Figure 3.1 - Relation between Arias Intensity Ia and significant duration D5-75 of one ground
motion record

To isolate the effects of duration, the selection of ground motions follows approaches
similar to the ones used in Barbosa et al. (2017) and Chandramohan et al. (2016) for developing
2-D structural analyses are followed. However, a new procedure is proposed to be used in 3-D
structural analyses, in which the spectral shape of bi-directional components of short-duration
ground motions are matched with bi-directional components of long-duration ground motions.
When analyzing 3-D structural models subjected to two orthogonal components of a ground
motion, the ground motion selection needs to carefully consider spectral matching in two
orthogonal directions to minimize effects of ground motion selection and interaction effects
between ground motion intensity and ground motion duration. This selection procedure
minimizes the root mean square error between the two pairs of rotated ground motion
components (for short and long-duration). To assess possible different ground motion directional
effects on the plan asymmetric building, both long- and short-duration ground motions are
applied in two orthogonal directions. To track damage, the Park and Ang (1985) and Reinhorn
41

and Valles (1996) damage indices are used, and based on them, damage index-based fragility
curves are proposed. The results in this study indicate that ground motion duration plays an
important role in the damage state of buildings.

3.3 Damage Indices

The structural damage can be quantified using damage indices that are function of
deformation, ductility, stiffness degradation, and energy dissipated, among others. Damage
indices available in the literature can be classified as (1) deformation-based, (2) energy-based, or
a combination of both, i.e., (3) deformation- and energy-based indices. For the deformation-
based damage indices, Veletsos and Newmark (1960) first proposed a damage index based on
the peak inelastic deformation. Roufaiel and Meyer (1987) proposed global damage measures
based on peak roof displacement and roof displacement at failure of reinforced concrete moment
frames. Banon et al. (1981) and Lybas and Sozen (1977) developed damage indices in which the
concepts of cyclic deterioration were incorporated in the analysis of structural systems consisting
of reinforced concrete moment frames and reinforced concrete coupled walls, respectively.
Stephens and Yao (1987) developed a cumulative damage index based on displacement ductility,
after testing two concrete structures and studying the correlation between the damage observed
in both. Jeong and Iwan (1988) proposed a classical low-cycle fatigue formulation calibrating
their model with observed failures of reinforced concrete columns. Chung et al. (1987) extended
the concept of damage and structural failures and developed a damage index formulation that
also accounts for fatigue. In terms of energy based damage indices, Gosain et al. (1977)
developed a cumulative energy ratio for reinforced concrete members. Darwin and Nmai (1986)
applied a factor to the damage index developed by Gosain et al. in order to account for the
reinforcing steel layout of reinforced concrete beams. Kratzig et al. (1989) proposed a more
complex energy formulation based on the definition of primary and follower half-cycles for
reinforced concrete members. To quantify the structural damage, McCabe and Hall (1989)
considers the energy dissipated as the main component in the evaluation of two damage
measures, one based on the concept of equivalent hysteretic cycles of deformation and the
second on concepts derived from low‐cycle fatigue theory. Fajfar (1992) and Powel and
Allahadabi (1988) also used the number of cycles, as well as the amplitude or history of the
42

inelastic cyclic response, to define two other damage indices. Regarding deformation- and
energy-based combined damage indices, Bracci et al. (1989) proposed a combined approach that
defined the potential damage as the total area between the monotonic load-deformation and the
fatigue failure envelope. DiPasquale and Cakmak (1990) presented yet another procedure to
evaluate the damage in the structure that is related to the overall stiffness loss in the structure due
to inelastic response through tracking the elongation of the fundamental period of vibration. Park
and Ang (1985) proposed an index based on the combination between maximum deformations in
the structure and the effect of the repetition of load cycles thought a term that considers the total
energy dissipated and the energy dissipated during the loading cycles. Reinhorn and Valles
(1996) extended the Park and Ang index to account for a damage measure that was calibrated
taking into consideration an empirically calibrated fatigue-based damage model.

From all the cumulative damage indices, the Park and Ang damage index (DIP&A) is the best
known and widely used and it is thus chosen to be applied in this study. In addition, the Reinhorn
and Valles damage index (DIR&V) is also here considered since it corresponds to an extension of
the Park and Ang that does consider the effect of low-cycle fatigue which could be important
when considering long-duration motions. The formulations for the DIP&A and DIR&V are hence
described in the following. The DIP&A is given by:

m 
 u  u Fy  h
DI P&A   dE (3.1)

and the DIR&V is set by:

m   y 1
DI R&V  (3.2)
u   y 
1   dEh 
 4(  u   y ) Fy 
 

where δm corresponds to the maximum deformation obtained, δu is the ultimate deformation


capacity, Fy is the yield strength of the element; and  dEh is the energy dissipated by yielding of

the element. Specifically, for the DIP&A, the parameter β is a non-negative factor which also
helps to define the capacity of the structure/element. In turn, DIR&V makes use of δy, which is the
deformation at a reference yield point.
43

The damage indices can be obtained for element sections (hinge damage index), for an
element (maximum of damage indices observed at all hinges in an element), per story, or for the
complete (global) structure. The following expressions are used to develop the section damage
indices based on the Park and Ang index:
m 
u u M y  h
DI P&A,section   dE (3.3)

and based on the Reinhorn and Valles damage index:

m   y 1
DI R&V ,section  (3.4)
u   y 
1   dEh 
 4  u   y  M y 
 

where the reference yielding rotation and reference yielding moment, θy and My, respectively, are
estimated using expressions in Priestley et al. (2007); ultimate rotations, θu, are here defined
using formulations proposed in ATC 72 (Malley et al. 2010), which correspond to 1.5 times
capping rotation (rotation at peak strength capacity) proposed by Haselton et al. (2008); β is a
function of the shear span ratio, axial force, longitudinal steel and confinement, and it is typically
defined as 0.05 for RC components, and is a parameter estimated from over 400 tests to
reinforced concrete beams and columns of rectangular cross section (Park et al. 1987; Fajfar
1992). While, the damage indices shown in Equations (3.1) and (3.2) were developed to quantify
the damage at the element level, to obtain the story and global damage index, the following
expressions are used for both Park and Ang and Reinhorn and Valles indices:
DI story j    i element  DIi element (3.5)

DI global     j   DI 
j story (3.6)
story

where:
 
 E 
 i element   ne i  (3.7)
 
  Ei 
 i 1 element
44

 
 E 
 
j story   ns

j 

(3.8)
  Ej 
 j 1  story

and where ne is the number of elements in story j, ns is the total number of stories, (λi)element
represents the ratio of yielding energy dissipated by one element to the total yielding energy
dissipated by the whole story, and (λj)story the ratio of yielding energy dissipated by respective
story to the yielding energy dissipated by the whole building.
It is worth noting that the global damage index may not be a useful measure to define the
damage state of the building because it does not provide an indication of the global failure
mechanism, such as in the case of a building with a soft story mechanism.
In term of defining damage states based on the values of the damage indices, Park et al.
(1985) first suggested a damage index value of 0.4 and above to represent severe damage, while
values of 1.0 or greater were defined to correspond to building collapse. Later Park et al. (1987)
further calibrated the damage index proposed by Park and Ang (1985) with damage observed in
nine reinforced concrete buildings damaged by the 1971 San Fernando and 1978 Miyagiken-Oki
earthquakes. The proposed breakdown for classifying damage states based on this work is shown
in Table 3.1. The same breakdown was later also adopted for the DIR&V damage index. Stone and
Taylor (1993) proposed new limits for the damage states based on the Park and Ang damage
index for use for circular bridge columns, and Ang et al. (1993) proposed further adjustments for
use in bridge structures. Table 3.1 summarizes the damage states and corresponding values of the
damage indices as suggested by the referenced authors.
45

Table 3.1 – Damage Index Classification

Damage Index
Park et al. Ang et al. (1993) Stone and Other
Damage State
(1987) Taylor (1993) observations
RC Buildings Bridge Structures Bridge Columns
Minor damage < 0.25 < 0.25 < 0.11 Minor Cracking
Moderate damage 0.25 - 0.4 0.25- 0.4 0.11 - 0.4 Repairable
Severe damage 0.4 - 1.0 0.4 - 0.8 0.4 - 0.77 Beyond repair
Collapse ≥ 1.0 ≥ 0.8 ≥ 0.77 Complete loss

The damage indices in each hinge are calculated separately for X and Y directions and then
the Square-Root-of-Sum-of-Squares SRSS of the values of damage indices for X and Y are given
by:

 m,X 
u u M y ,X  h,X
DI P&A,section,X   dE (3.9)

 m,Y 
u u M y ,Y  h,Y
DI P&A,section,Y   dE (3.10)

DI P&A,section  DI P&A,section,X 2  DI P&A,section,Y 2 (3.11)

Although the equations above show the described methodology applied only to DIP&A, it is
important to note that the DIR&V follows similar equations.

3.4 Ground motion selection

This paper analyses a 3-D structural model subjected to orthogonal horizontal ground
motion components of the earthquake records, and therefore matching of the response spectra of
long-duration and short-duration ground motion records for each component becomes a non-
trivial task, when isolating the effects of duration. It is important that the ground motion
selection minimizes the disparity between the response spectra of long-duration and short-
duration earthquake records for both components. To address this issue, the methodologies in
Barbosa et al. (2017) and Chandramohan et al. (2016), developed for 2-D structural analyses, are
extended for application to 3-D analyses. In this study, the response spectra of two orthogonal
46

components of rotated short-duration ground motion components are matched to as-recorded


long-duration ground motion components. Rotation of the short-duration ground motion
components allows for the minimization of differences in the response spectra of short- and long-
duration ground motion pairs, thereby providing for an improved way for isolating the effects of
ground motion duration for 3-D analyses. The methodology for selecting the long-duration
ground motion and the short-duration ground motion sets is described in the next two
subsections. It is worth noting that for this work, an earthquake is classified as a long-duration
earthquake when at least one of its orthogonal components has a significant duration (D5-75)
greater than 25 seconds (Chandramohan et al. 2016).

3.4.1 Long-duration ground motion set


The long-duration earthquake records were selected from a set that included the earthquakes
of Valparaiso and Maule in Chile (1985 and 2010, respectively), Tohoku in Japan (2011), and
long-duration earthquake records from PEER NGA-west2 database (Ancheta et al. 2014). The
compiled database of records includes 32 long-duration ground motion pairs described in Table
A1 (in Appendix). The Chilean earthquake records were obtained from COSMOS Virtual Data
Center (COSMOS 2012) and Center for Engineering Strong Motion Data (Haddadi et al. 2008),
and the ones of Tohoku were taken from K-NET and Kik-net database (National Research
Institute for Earth Science ans Disaster 1996). All long-duration ground motions, with exception
to ones from the PEER database, were filtered and baseline corrected using the recommendations
of Boore and Bommer (Boore and Bommer 2005; Boore 2005). More details on this work can be
found in Long (2012).

3.4.2 Short-duration ground motion set


The algorithm proposed for selection of the short-duration records is:
Step 1: Select a source for the short-duration earthquake records, for example, the PEER NGA
ground motion database. Short-duration earthquake records are characterized here as having
both horizontal components with significant durations less than 25 seconds.
Step 2: For the short-duration earthquake records, rotate the as-recorded pairs of orthogonal
horizontal motions through possible non-redundant rotation angles (θj with increments of 10
degrees, for example), compute the response spectra for each rotated component, and then
47

compute geometric mean of the response spectrum for the rotated components for each used
rotation angle.
Step 3: For each pair of the short-duration earthquake records of θj rotated records, select a single
scaling factor SF that yields the minimum RMSEGM, which is given by:

min RMSEGM  j   W T   ln  Sa T    ln  SF  Sa T ,  W T   ln  Sa T    ln  SF  Sa T , 


2 2
i
long
H1 i
short
H1 i j  i
long
H2 i
short
H2 i j
 j i i

(3.12)
where W(Ti) is a modulation indicator function taking a value of 1.0 for each discretized
periods Ti falling within a period range of interest (typically 0.2T1 < T1 < 3.0T1, where T1 is
the fundamental period of the structures), and zero for values of Ti elsewhere; SaH1 Ti  is
long

the 5-percent damped linear elastic spectral acceleration of the first component (EW
direction) of the unscaled long-duration ground motion record and SaH2 Ti  is 5-percent
long

damped linear elastic spectral acceleration of the unscaled long-duration ground motion
record for an orthogonal component (NS direction); SaHshort Ti , j  and SaHshort Ti , j  are 5-
1 2

percent damped linear elastic spectral accelerations of two orthogonal components H1 and
H2 of a short-duration ground motion rotated by θj degrees.
The difference of spectral accelerations in equation (3.12) is taken in natural log-space,
since ground motion spectral accelerations have been shown to follow a lognormal
distribution (Abrahamson 1988; Baker and Cornell 2005; Baker 2015), where the spectral
accelerations of the rotated short-duration motions are given by:

SaHshort
1
Ti , j   SaHshort
1
Ti   cos  j   SaHshort
2
Ti   sin  j  (3.13)

SaHshort
2
Ti , j   SaHshort
2
Ti   cos  j   SaHshort
1
Ti   sin  j  (3.14)

Step 4. The short-duration ground motions selected correspond to ones producing the
smallest value of the short-duration earthquakes minimum RMSEGM obtained in step
3.
It is important to note that all ground motions classified as pulse-like ground motions (Baker
2007) were not included in these analyses. In addition, also worth noting, the period range of
48

interest selected in this work is 0.2T1 < T1 < 3.0T1, and thus the modulation function takes values
of 1.0 in this range.

3.4.3 Ground Motions Characterization


Table A1 (in Appendix) lists the pairs of long-duration and short-duration of earthquake
records selected. In addition, the scale factors and rotation angles applied to each short-duration
earthquake record are also shown. Tables A2 and A3 (in Appendix) characterize the components
of the selected long-duration and scaled short-duration earthquake records in terms of D5-75,
rupture distance, peak ground acceleration (PGA), pulse index (PI), pulse period (Tp), magnitude
(MW), and the 5-percent damped linear elastic spectral acceleration at the fundamental period of
the structure T1. Figure 3.2 shows the response spectra for an example of a pair (pair #1 in Table
A1) of the selected earthquake records using the procedures described in section 3.4.1 and 3.4.2,
while Figure 3.3 shows the respective time-histories for the same earthquake record pair. Figure
3.4 displays the median acceleration response spectra of components H1 and H2 as well as the
16th and 84th percentiles spectra. For reference, the bounds of the period range considered in the
short-duration ground motions selection (0.2T1 and 3.0T1), Sa(T1), and Sa(1.5T1) are identified. It
can be seen that there is an excellent agreement between the response spectra for the short- and
long-duration components, not only in terms of median response spectra, but also in terms of the
16th and 84th percentiles.
49

(a) (b)

(c)

Figure 3.2 – Response spectra for earthquake record pair #1, including the Tohoku (2011)
Towadako long-duration earthquake record and the scaled and rotated Hector Mine (1999)
Hector short-duration earthquake record: (a) Horizontal component H1; (b) Horizontal
component H2, and (c) Geometric mean of the two horizontal components
50

(a)

(b)

(c)

(d)
Figure 3.3 – Acceleration time-series for earthquake records of pair #1: (a) Tohoku (2011)
Towadako station, horizontal component H1; (b) Hector Mine (1999) Hector station, horizontal
component H1; (c) Tohoku (2011) Towadako station, horizontal component H2; and (d) Hector
Mine (1999) Hector station, horizontal component H2.
51

(a) (b)

Figure 3.4 – Response spectra for the ground motion set selected: (a) Horizontal component
H1; (b) Horizontal component H2

Figure 3.5 (a) shows the distribution of the significant duration D5-75 for all short- and long-
duration ground motion components, where it can be seen that the mode of the significant
duration is less than 10 sec for the short-duration records and between 30 and 40 sec for long-
duration records. Figure 3.5 (b) shows that the ratios between the significant durations of the
long-duration horizontal ground motion component and the corresponding short-duration motion
component is less than 5 for more than half of the cases. Figure 3.5 (c) shows the scale factors
applied to the short-duration ground motions in order to match the long-duration motions, with
most of the scale factors falling between 0.5 and 2.

Figure 3.6 displays the correlation between significant duration and the natural logarithm of
the 5-percent damped linear elastic spectral accelerations versus the period of vibration; it can be
observed that the correlation coefficients are below 0.2 for the period range of interest, which is
important so that the duration effect can be isolated in the subsequent analyses.
52

(a) (b)

(c)
Figure 3.5 – Histograms for ground motions components for all earthquake records used in this
study: (a) Significant duration D5-75; (b) ratio of D5-75 of long-duration motions to D5-75 of short-
duration motions; and (c) scale factor applied to short-duration earthquake records.
53

Figure 3.6 – Correlation coefficient between significant duration, D5-75, and natural logarithm of
the spectral acceleration versus period of vibration.

3.5 Building case study

3.5.1 General Description


The case study building used in this study is a 3-story building known as SPEAR building
that represents typical old non-ductile concrete moment frame buildings. In its design, only
gravity loads were considered based on the construction practice applied in the decade of 70.
This building case study was selected because it is of interest not only for Southern Europe,
where the prototype structure originates from, but also for other regions in the world that have
old non-ductile reinforced concrete structures.

A full-scale specimen of the SPEAR building was tested under pseudo-dynamic conditions,
subjected to a bi-direction seismic loading as part of the European SPEAR project (Fardis and
Negro 2005).

Figure 3.7 shows the SPEAR building experiment specimen in the ESLA laboratory in
Europe. The building is plan-asymmetric in both X and Y directions and regular in elevation as
can be seen in Figure 3.8 (a) and Figure 3.8 (b). Eight of the nine columns have a 0.25×0.25 m2
54

cross-section reinforced with four (4) 12mm longitudinal plain bars and column C6 has a
0.25×0.75 m2 cross-section reinforced with ten (10) 12mm plain bars as shown in Figure 3.8 (c).
Figure 3.8 (d) shows a typical beam cross–section close to a support. All beams are rectangular
with 0.25×0.50 m2 with two 12mm plain bars (top and bottom) at mid-span and four 12 mm
plain bars on top over supports. Beams B4, B7, and B9, have different detailing, with 20 mm
reinforcing steel bars (Fardis and Negro 2005). In both columns and beams, the stirrups are 8
mm plain bars with 250mm spacing. The concrete material used in the building has compressive
strength of 25 MPa, whereas the reinforcing steel has yield strength of 360 MPa. The weight of
the building is distributed with a total mass of 67.3 tons in first and second floor levels and 62.8
tons in the third floor. Further details on the 3-story building characteristics can be found in
(Fardis and Negro 2005).

Figure 3.7 – SPEAR building specimen (Fardis and Negro 2005)


55

(a) (b)

(c) (d)
Figure 3.8 – Geometry and detailing of the Spear 3-story building: (a) Plan view; (b) Elevation
view; (c) Column cross sections; (d) Typical interior beam cross-section. All dimensions are in
meters

3.6 Nonlinear Finite Element Model Development and analysis

A nonlinear finite element model of the building was developed in OpenSees (McKenna et
al. 2010). The material models used in the fiber-section discretization were: (1) a uniaxial
Popovics concrete material (Popovics 1973) with degraded linear unloading/reloading stiffness
according to the work of Karsan and Jirsa (1969); tensile strength with exponential tension
stiffening (Concrete04 material in OpenSees); (2) the Giuffre-Menegotto-Pinto (Menegoto and
Pinto 1973) as modified by Filippou et al. (1983) (Steel02 material in OpenSees) was assigned to
the reinforcing steel bars.

In the nonlinear finite element model, beams and columns are modelled using fiber-section
displacement-based nonlinear frame elements. Each column or beam are discretized in four (4)
56

or five (5) elements with three (3) Gauss–Legendre integration points (i.e., monitored cross
sections) along each element length. At each integration point, the section is discretized using a
fiber model, comprised of approximately 100 fibers. The cross-sectional response can capture the
stiffness degradation and strength deterioration due to concrete cracking, concrete crushing, and
steel yielding, based on the uniaxial material constitutive laws adopted. The fiber cross-section
automatically accounts for the interaction between axial force and biaxial bending. Element
shear as well as torsional behaviors are assumed linear elastic, uncoupled, and aggregated to the
nonlinear inelastic fiber cross-section behavior. Thus, the frame elements do not account for
stiffness degradation and strength deterioration in the shear and torsional behavior of the beams
and columns. The length of the elements for columns and beams correspond to the estimated
plastic hinge length lp, of the respective member using the empirically validated relationship
proposed by Paulay and Priestley (1992). The plastic hinge length lp is measured from the axis of
the connected member, implying that the effective plastic hinge length, measured from the face
of the connecting element, is reduced, to indirectly capture the effect of using plain bars, at the
global level of response. The columns were modelled using as-built dimensions from the test and
beam flange widths were included in the model to capture the effects of the slab flexural stiffness
according to reinforced concrete provisions (EN 1992-1-1 2004). In the in-plane direction, floors
were modelled as rigid diaphragms, which was a reasonable option for this building that
experienced little to no damage in the beams during testing. Gravity loads and masses of each
floor were applied at the beam-column joint nodes and assigned based on their tributary areas. A
damping ratio of 2 percent was assigned to the building model at the fundamental period of the
structure and at a period of 0.2 times the fundamental period, assuming a Rayleigh-type damping
proportional to the mass and tangent stiffness. The damping ratio values used are based on
experimental results (Negro et al. 2004; Fardis and Negro 2005). In terms of time history
response analyses, the Newmark average acceleration method was employed. The Newton-
Raphson method is used to solve the nonlinear system of dynamic equations of equilibrium at
each time step.

3.6.1 Model Validation


The bi-directional seismic loading applied during the experimental tests on the SPEAR
building was the ground motion recorded in Hercegnovi station during the Montenegro 1979
57

earthquake (a short-duration ground motion) matched to Eurocode 8 (EN 1998-1 2004) spectrum
(type 1, soil C). The test structure was subjected to three motions corresponding to three different
intensities with PGA values of to 0.02g, 0.15g, and 0.2g.

The finite element model developed in this study was validated through the comparison of
periods of vibration and dynamic response using the available experimental results obtained for a
PGA of 0.2g. Table 3.2 shows the modal properties of the analytical model, after application of
gravity loads. It is worth noting that these were not measured for the building specimen. Instead,
the periods measured for the experiment were estimated from the time-history response with
PGA 0.02g (Negro et al. 2004; Fardis and Negro 2005). The values estimated in the experiment
are 0.84 sec, 0.78 sec, and 0.69 sec, for the first three modes respectively. One should note that
there is a discrepancy between the very initial periods of the finite element model and the periods
of vibration measured in the test. Note that the discrepancy is part due to fact that the periods
from the test were measured after the structure had been moved to the test location and in this
process the structure was partially damaged (cracked) (Negro et al. 2004; Fardis and Negro
2005). However, this process was not simulated in the OpenSees model developed.

Table 3.2 – Numerical model and experiment specimen modal properties of the 3-Story
building

Periods Effective Modal Mass


Mode of vibration,
UX UY RZ
T
1 0.69 61% 8% 19%
2 0.59 23% 42% 20%
3 0.49 3% 32% 49%
4 0.25 7% 1% 2%
5 0.20 3% 4% 4%
6 0.16 2% 0% 1%
7 0.16 0% 8% 4%

Figure 3.9 shows time history results obtained in the experimental and the analytical model,
for the structure subjected to the Hercegnovi records for a PGA of 0.2g. It is worth noting that
the actual sequence of motions experienced by the test structures were used as input, i.e. the
0.02g, 0.15g and 0.2g scaled motions in sequence, although results are here shown for the 0.2g
case only. In this figure, four responses are shown: (a) the X-component of the roof displacement
58

at the location of column C1, (b) the Y-component of the roof displacement at the location of
column C9, (c) the X-component of the first story IDR of column C1, and (4) the Y-component
of the second story IDR of column C9. From the observation of these results, it is worth
mentioning that a reasonable approximation is obtained between experimental and analytical
results, both in terms of peak deformations and number of cycles. Even though it is clear that
some discrepancies exist between the experimental and numerical results, some of the
simplifications in the modeling both in the sequence of loading applied to the models (e.g.,
effects of moving of the specimen to its testing location, loading sequence prior to the 0.2g PGA
testing sequence) and also in terms of model simplifications (e.g., bond-slip is not explicitly
modelled, although plastic hinge lengths were defined to localize damage near the ends; and,
beam-column joints not explicitly modeled) can be responsible for some of the discrepancies
observed. In the interest of the main objective of the paper, which is to study the effect of ground
motion duration, however, the fact that peak responses and the number of cycles are similar in
the experimental and numerical results, indicate that the nonlinear finite element model is
acceptable for the ground motion duration study performed herein.

To further validate the finite element model and the use of the damage indices and their
values (Table 3.1) for this RC building, the damage states of the specimen after the test (Negro et
al. 2004) were analyzed and compared to the damage indices DIP&A and DIR&V computed from
the numerical results for base of column C3, top of column C1 at the second story, and top of
column C3 at the second story, as shown in Figure 3.10 (a), (b) and (c), respectively. It can be
seen in the Figure 3.10 that the observed damage matches the values of damage indices that fall
within the ranges of values proposed by Park et al. (1987) used in the DIP&A and DIR&V.
59

(a) (b)

(c) (d)

Figure 3.9 – Correlation between the experimental and analytical results: (a) Roof displacement
in X direction at column C1; (b) Roof displacement in Y direction at column C9; (c) Interstory
drift in X direction at column C1 – 1st story; (d) Interstory drift in X direction at column C9 -
2nd story. Experiment and analytical model developed using OpenSees were subjected to the
Hercegnovi record from the Montenegro (1979) earthquake scaled to a PGA of 0.2g.
60

(a) (b)

(c)
Figure 3.10 – Observed damage on members after experiment (Negro et al. 2004) with computed
damage indices from numerical analysis: (a) Damage state of base of column C3 - Minor
damage; (b) Damage state of top of column C1 at 2nd story - Moderate damage; (c) Damage state
of top of column C3 at 2nd story - Severe damage

3.7 Results and Discussion

3.7.1 Global response


Structural deformation, together with forces and dissipated energy are the main response
measures used to quantify structural damage in buildings. Furthermore, these are the response
measures that play main role when calculating the damage indices used in this study, including
the DIP&A and DIR&V described in Section 3.3. Incremental Dynamic Analyses (IDA) were
61

performed in which the engineering response parameters (deformation, accelerations, forces,


energy dissipated) were quantified using a nonlinear finite element of the building structure.
Figure 3.11 shows the IDA curves for the magnitude of vector sum of the peak IDR in X- and Y-
directions when subjected to the two sets of 32 long-duration, and scaled and rotated short-
duration records. The vector sum consists in the square root of the sum of the squares (SRSS).
The figure shows IDA curves of both sets of ground motions applied to the building as well as
the median IDA curves for each ground motion set. It can be seen from Figure 3.11 that the
median IDA curves obtained from short- and long-duration ground motions are similar and only
a slight difference can be noted in terms of the variability of results, in which short-duration
ground motion results show slightly greater deviations. It is worth noting, however, that based on
the IDA curves alone, the effect of duration on IDR may not be sufficient for assessing this
effect. In fact, only detailed analysis of the responses could rule out the effect of ground motion
duration, as illustrated in the results shown for the damage indices as a function of the increasing
shaking intensity discussed below.

Figure 3.11 – Incremental dynamic analysis curves for peak of the vector sum of interstory
drift ratios in X- and Y-directions.
62

Figure 3.12 shows median IDRs envelope measured at the location of column C8 in the X-
direction, C2 in Y-direction, and the vector sum of the IDR in two orthogonal directions at the
location of column C3, for two levels of shaking intensity, Sa(T1) = 0.2g (Figure 3.12 a to Figure
3.12 c) and 0.4g (Figure 3.12 d to Figure 3.12 f). Note that Figure 3.12 (d) to Figure 3.12 (f)
exhibit the results obtained from all analyses with no filter applied and Figure 3.12 (g) to Figure
3.12 (i) shows only non-collapse cases. The term “non-collapse case” corresponds to a time
history response analysis in which the algorithms used for solving the nonlinear system of
equations converged for the duration of the applied ground motion and peak IDRs is less than 10
percent. Thus, a “collapse case” is a case in which the analysis showed lack of convergence
accompanied by identified structural instabilities, or when the peak IDR exceeded 10 percent
drift. In all these cases, the damage indices were computed and checked to be greater than 1.0.
63

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 3.12 – Median interstory drift ratio envelopes: (a) C8 in X direction for, considering
all cases; (b) C2 in Y direction for Sa(T1) = 0.2g, considering all cases; (c) C3 Vector Sum for
Sa(T1) = 0.2g, considering all cases; d) C8 in X direction for Sa(T1) = 0.4g, considering all cases;
(e) C2 in Y direction for Sa(T1) = 0.4g, considering all cases; (f) C3 Vector Sum for Sa(T1) =
0.4g, considering all cases; (g) C8 in X direction for, considering only the non-collapsed cases;
(h) C2 in Y direction for Sa(T1) = 0.4g, considering only the non-collapsed cases and; (i) C3
Vector Sum for Sa(T1) = 0.4g, considering only the non-collapsed cases
64

(a) (b)
Figure 3.13 – Statistics of normalized absolute roof displacements u at three column locations
(C8, C3, C2) relative to the absolute roof displacement of the centre of stiffness uCR: (a) instant
when peak u/uCR occurs at C2; (b) instant when peak u/uCR occurs at C8.

Comparison of these two sets of figures (Figure 3.12 d to Figure 3.12 f with Figure 3.12 g to
Figure 3.12 i) highlights that the main mode of collapse is a soft-story mechanism at the first
story. Comparison of the results between both sets of ground motion records indicates that the
ground motion duration does not have a noticeable effect on the median IDR envelopes.

Figure 3.13 shows the torsion rotation of the building through normalization of roof
displacements in columns C2 and C8 relative to column C3. Figure 3.13 (a) and Figure 3.13 (b)
show normalized displacements at instants in which peak normalized roof displacements in C8
and C2 are obtained, respectively. The torsional rotation of the building is strongly evident when
the building behaves mainly in the elastic regime (Figure 3.13 a). For larger intensities, the
building still exhibits torsional deformations, but these are not as pronounced. Despite small
differences in the results obtained between short and long-duration, results indicate that duration
does not seem to affect the building displacement response.
65

Although not shown in this figure, it is also worth noting that analysis of the drift response
for all columns indicated that the columns that belong to the same alignments in X direction have
similar peak IDRs in X direction and columns in the same alignments in Y direction also have
similar peak IDRs in the Y direction. It was also observed, the alignments C5-C1-C2 and C7-C4-
C2 correspond to the most flexible column alignment in the building in X and Y directions,
respectively, caused by torsion of the building.

3.7.2 Damage Assessment


In order to assess the damage in the building, both peak deformation responses and
cumulative response measures should be considered. For the moment frame building model, in
which nonlinearity can only occur in beams and columns, the yielding energy dissipated in the
X-direction, Y- direction, and the magnitude of the vector sum of the dissipated energy, EY , are
given by, respectively:

c by
E   M i    M j j
Y
y x
i
x
(3.15)
i 1 j 1

c bx
EYx   M iyiy   M j j (3.16)
i 1 j 1

EY  EYy 2  EYx 2 (3.17)

where c represents the number of columns, bx and by are all beams oriented in the X-
x y
direction and Y-direction respectively, M i and M i are the bending moment in the column end

sections about X and Y building global axis, respectively, and Mj corresponds to the bending
moment in the beam end sections. Figure 3.14 graphs the total yielding energy dissipated versus
Sa(T1). As expected, it can be seen that considerable larger amounts of yielding energy is
dissipated when the structure is subjected to long-duration ground motions. For the ground
motion sets used, the increase the ratio of the peak median yielding energy dissipated due to
long-duration motions to the peak median yielding energy dissipated due to the short-duration
motions is approximately 4.0, matching with the ratio between significant durations of long-
duration to short-duration ground motions that shows the mode value in the 2.5-5.0 range.
66

Figure 3.14 – Sum of yielding energy dissipated in beam and column hinges for short- and long-
duration ground motion sets.

The yielding energy dissipated plays an important contribution in the determination of the
damage indices by Park and Ang and Reinhorn and Valles. These indices can be obtained for
element, story, and whole structure using the methodology described in Section 3.3.

Figure 3.15 and Figure 3.16 show, respectively, the DIP&A and DIR&V associated to columns
and beams in alignment C5-C1-C2 of the building for a set of different intensities, as well as the
referred damage indices at the column end sections (plastic hinges) for a relatively large ground
motion intensity, Sa(T1) = 0.5g. From observation of both Figure 3.15 (a) and Figure 3.15 (c), it
can be inferred that the collapse mechanism of the structure corresponds mainly to a soft-story at
the first story. On the other hand, the damage developed in the beams is negligible as observed in
Figure 3.15 (b) and Figure 3.16 (b). For values of Sa(T1) = 0.5g the values of DIP&A obtained for
the columns in first story (Figure 3.15 a) show that they reached collapse when considering both
short- and long-duration ground motions. Regarding to the DIR&V, the effect of ground motion
duration is noticeable when observing Figure 3.16 (a), which in the results indicate that the
columns in the first story are closer to reach collapse for long-duration ground motions. For both
damage indices obtained for the first story columns, it can be observed that the difference in the
67

results is significant in the transition from Sa(T1) = 0.4g to Sa(T1) = 0.5g. When analyzing the
second story, this difference is not so sharp and, in general both damage indices show greater
results for long-duration ground motions. Here the term ‘collapse’ intends to represent physical
meaning to building depending on the value of damage index, according to Park et al. (1987).
68

(a) (b)

(c)

Figure 3.15 – Park and Ang damage index results for structural alignment C5-C1-C2: (a) Median
damage values for columns at five intensities of shaking (0.1g, 0.2g, 0.3g, 0.4g, and 0.5g); (b)
Median damage values for beams at five intensities of shaking (0.1g, 0.2g, 0.3g, 0.4g, and 0.5g);
(c) Damage index for hinges at column ends for the Sa(T1) = 0.5g intensity level.
69

(a) (b)

Sa(T1) = 0.5g

(c)

Figure 3.16 – Reinhorn and Valles damage index results for structural alignment C5-C1-C2: (a)
Median damage values for columns at five intensities of shaking (0.1g, 0.2g, 0.3g, 0.4g, and
0.5g); (b) Median damage values for beams at five intensities of shaking (0.1g, 0.2g, 0.3g, 0.4g
and 0.5g); (c) Damage index for hinges at column ends for the Sa(T1) = 0.5g intensity level

The global and stories damage indices versus Sa(T1) are displayed in Figure 3.17 (a) and
Figure 3.17 (b) for the DIP&A and DIR&V, respectively. Figure 3.18 (a) and Figure 3.18 (b) show
the envelope curves, i.e. an unique curve that contains the maximum value of damage index,
considering the damage indices for the three stories and global structure represented in Figure
3.17 (a) and Figure 3.17 (b). From this set of figures, it can be seen that both damage indices
70

identify that most of the damage is observed in the second story, until the moderate damage (DI
> 0.25) is reached. Beyond the moderate damage state, the first story is the one contributing most
the damage in the structure. For short-duration ground motions the same conclusion is taken
when considering the DIP&A, however the global DIR&V is the one that shows higher values after
moderate collapse is reached in the structure. When comparing the damage predicted by the two
damage indices, the DIR&V cases reach the severe damage (DI > 0.4), threshold for lower values
of Sa(T1) than the DIP&A cases. However, the collapse threshold (DI > 1.0) is reached for the
same intensity regarding both short- and long- duration ground motions.

(a) (b)

Figure 3.17 – (a) Global median DIP&A and story median DIP&A; (b) Global median DIR&V and
story median DIR&V
71

(a) (b)
Figure 3.18 – (a) Envelope for median DIP&A; (b) Envelope for median DIR&V

With respect to the duration effects, from observation of Figure 3.17 and Figure 3.18, it can
be seen that there is a negligible difference between damage identified using the DIP&A index for
short-duration and long-duration motions; in the case of the DIR&V, although noticeable,
differences are mainly observed at larger intensities of shaking.

3.7.3 Fragility Curve Development


With the damage indices obtained from all analyses, fragility curves that give the probability
of exceedance of a determined damage state were developed according with the damage
classification shown in Table 3.1. Figure 3.19 displays the probability of equaling or exceeding a
moderate damage state (DI > 0.25), a severe damage state (DI > 0.4), and collapse damage state
(DI > 1.0) obtained using both damage indices considered in this study, which correspond to
moderate damage, severe damage, and collapse fragility functions.

The development of the fragility curves is based on the computation of probabilities of


reaching a specified damage state, for all intensity levels considered, through calculation of the
ratio between cases where the threshold value corresponding to the damage state specified is
reached and the total number of cases analyzed.

The fragility functions are shown for the building at a global level, and also for each story.
From analysis of the fragility functions obtained based on use of DIP&A (Figure 3.19 a to Figure
72

3.19 c) one can say that duration does not affect the results significantly, except for the highest
levels of damage (above 0.6). Similar conclusions can be reached though observation of Figure
3.19 (d) to Figure 3.19 (f), for the fragility curves based on DIR&V.

Figure 3.20 (a) and Figure 3.20 (b) show probability of reaching severe damage state
comparing fragility functions obtained based on IDR, DIP&A and DIR&V for short- and long-
duration respectively. Following FEMA 356, an IDR of four percent was herein considered to
represent severe damage (ASCE 2000) and the fragility functions that show the greatest
probabilities, amongst the ones displayed in Figure 3.19 (1st to 3rd story and global damage
indices) are considered to characterize the structural damage.
73

1.0
Moderate Severe Collapse
damage damage damage
0.8 state state state
P [ DIP&A > di | Sa(T1) ]

di = 0.25 di = 0.40 di = 1.00

0.6

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Spectral Acceleration Spectral Acceleration Spectral Acceleration
Sa(T1), g Sa(T1), g Sa(T1), g

(a) (b) (c)

1.0
Moderate Severe Collapse
damage damage damage
0.8 state state state
di = 0.25 di = 0.40 di = 1.00
P [ DIR&V > di | Sa(T1) ]

0.6

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Spectral Acceleration Spectral Acceleration Spectral Acceleration
Sa(T1), g Sa(T1), g Sa(T1), g

(c) (e) (f)

Short-duration – overall Long-duration – overall


st st
Short-duration – 1 Story Long-duration – 1 Story
nd nd
Short-duration – 2 Story Long-duration – 2 Story
rd rd
Short-duration – 3 Story Long-duration – 3 Story

Figure 3.19 – Story and global structure fragility curves: (a) Moderate damage state based
on use of Park and Ang damage index; (b) Severe damage state based on use of Park and Ang
damage index; (c) Collapse based on use of Park and Ang index; (d) Moderate damage state
based on use of Reinhorn and Valles damage index; (e) Severe damage state based on use of
Reinhorn and Valles damage index; (f) Collapse based on use of Reinhorn and Valles index
74

(a) (b)

(c) (d)

Figure 3.20 – Fragility curves based in peak interstory drift ratio IDR and on damage indices
DIP&A and DIR&V: (a) Probability of severe damage on the building subjected to short-duration
ground motions; (b) Probability of severe damage on the building subjected to long-duration
ground motions; (c) Probability of collapse of the building subjected to short-duration ground
motions; (d) Probability of collapse of the building subjected to long-duration ground motions
75

Fragility curves defined by lognormal cumulative distribution functions were also included
in Figure 3.20 to fit the empirical fragility curves obtained. The fitted lognormal functions are
given by:

 ln  x   
P[Damage State ]     (3.18)
  

as suggested by Baker (2015) where Φ(.) is the standard normal cumulative distribution function,
x is the value of Sa(T1) considered, and the parameters μ and β correspond, respectively, to the
mean and standard deviation of the normal distribution that represents the lnSa(T1):

1 n
  ln Sa T1 
n i 1 (3.19)

1 n   Sa T1   
2

   ln  
n  1 i 1     
(3.20)

where n is the number of cases considered. In Figure 20, values exceeding four percent of IDR
were assumed to represent severe damage, while values of damage exceeding 0.4 and 1.0 were
considered for the severe damage states and collapse damage states for damage index-based
fragility curves, respectively.

From analysis of Figure 3.20 (a), for short-duration ground motions, the probability of
reaching both severe damage based on the IDR match with the probability based on any damage
index for all range of Sa(T1). For long-duration ground motions, greater probabilities of reaching
severe damage, for the same intensity level, are obtained when considering the damage indices
applied in this work, as shown in Figure 3.20 (b), reflecting the impact of the ground motion
duration in the damage of the building. In terms of the probability of being in the collapse state,
fragility functions based on the damage indices are here proposed rather than IDR-based
functions. Figure 3.20 (c) and Figure 3.20 (d) show similar fragility curves based in DIR&V and
DIP&A for short- and long-duration ground motions respectively for probability of obtaining
values of damage index greater than 1.0.
76

3.8 Conclusions

Recent long-duration earthquakes originating in subduction zone earthquake such as the


ones experienced in the subduction zones (Chile 2010; Japan 2011; Nepal 2015) regions have
brought the topic of ground motion duration and its impact on building and bridge response to
the research limelight. The effect of long-duration motions on structural response could be
important for regions that have experienced and can experience again great M8.5+ earthquakes.
These regions include, among others, the Pacific Northwest regions of the United States and
Canada, Southern Europe and Northern Africa (in the boundaries between the African and the
Eurasian plate), as well as in Nepal and surrounding countries (in the boundary of the Indian
plate and the Eurasian plate boundaries). Even though some recent studies have shown that
longer duration ground motions may lead to greater collapse risk, an extensive assessment of the
influence of ground motion duration on structural damage is still lacking in the earthquake
engineering literature. As a result, existing seismic risk methodologies are typically based on
response and damage measures calibrated to short-duration earthquake ground motions. This
paper aims at partially bridging that knowledge gap by providing new results for existing vintage
reinforced concrete buildings. Since torsion plays an important role in the seismic behavior of
this plan-asymmetric building, new questions arise regarding to the influence of torsion in the
assessment of ground motion duration. In this paper, the effect of earthquake ground motion
duration was evaluated on the response of a 3-story non-ductile reinforced concrete moment
frame plan-asymmetric vintage building. A ground motion selection procedure was proposed for
isolating the effects of duration in 3-D structural analysis, in which the spectral shape of bi-
directional components of short-duration ground motions are matched with bi-directional
components of long-duration ground motions. Thus, the ground motion duration assessment was
performed using a set of 64 bi-directional long-duration ground motions and a spectrally
equivalent bi-directional set of scaled and rotated short-duration ground motions, applied in two
orthogonal directions was selected. The building model analyzed corresponds to a full-scale
building that was tested at the ELSA laboratories in Europe in January 2004, which was designed
to withstand gravity loads only. The building was built based on a prototype of common building
construction practice of Southern Europe prior to the 1980s, which are a prevalent form of many
existing buildings in seismically active regions of both the western US and in Southern Europe.
77

A nonlinear finite element model of the 3-story building was developed in OpenSees. The finite
element model was developed specifically with the intent of analyzing the ground motion
duration effects, and validation between experimental results and analytical results presented in
this work show a reasonable approximation in terms of peak responses as well as number of
cycles, which are important response parameters in this study. Results from the damage
assessment performed herein are based on the damage indices by Park and Ang (DIP&A) and by
Reinhorn and Valles (DIR&V). This study helps to understand how the ground motion duration
could impact the structural behavior of buildings. A full-scale 3-story building tested in Europe,
and representative of non-seismically designed structures existing in many parts of seismic areas
of the world, was here used as a case study in the development of the three-dimensional
numerical model. This work can be considered as a bridge between all work done when using
two-dimensional numerical models. New collapse fragility curves based on DIP&A and DIR&V are
proposed for this type of building to account for the effects of ground motion duration.

The results in this paper leading to the following conclusions:

 For the building studied, collapse was reached for median values of Sa(T1) between 0.4g
and 0.5g, which corresponds to a range of PGA values between 0.22g and 0.27g. A soft-
story mechanism in the first story of the building was the main mode of observed
building collapse.
 The duration does not play an important role in terms of peak displacement responses for
this non-ductile reinforced concrete building. In terms of torsional response of the plan-
asymmetric building analyzed herein, ground motion duration also had a negligible
influence on the observed drift responses.
 In terms of yielding energy dissipated, the long-duration motions induce greater energy
dissipation. A median value of 4.0 was observed for the ratio of between yielding energy
dissipated during long-duration motions and short-duration motions.
 Damage estimated using the Park and Ang and the Reinhorn and Valles damage index
show minor increase in damage due to duration effects.
78

 Fragility curves developed based on the Park and Ang and the Reinhorn and Valles
damage index damage indices show similar shapes, both for severe damage and collapse
limit states.
 When considering short-duration ground motions for severe damage state fragility
curves, IDR-based and damaged-index based fragility curves are identical. However,
when considering long-duration motions, there is a noticeable difference between the
IDR-based and damage index-based fragility curves at larger shaking intensities.

In summary, the results shown in this work show that the ground motion duration effects are
evident only for intensities that lead to collapse of the structure. However, the ground motion
duration did not play an important role on the displacement and drift responses of this 3-D plan
asymmetric RC building. The reason for that might be related to the brittle behavior of this
specific building, since there is not enough capacity in the structural members to reach large
deformations. Nonetheless, minor differences were noted in the local damage predicted in the
first story columns and second story columns, especially at larger shaking intensities.

Damage index-based fragility curves proposed in this study capture the effects of duration
for larger intensities of shaking, unlike the IDR-based fragility curves. Thus, the results indicate
that the damage index-based fragility curves should be developed for capturing the effect of
ground motion duration when performing a structural damage and loss assessment.

In terms of potential future work on the effects of ground motion duration of non-ductile
reinforced concrete buildings, the following points could be addressed:

 The impact of different modelling options such as the effect of bond slip, bar buckling,
strength and stiffness degradation of beam-column joints, or shear failures, could be
explored.
 Different damage indices could be considered (e.g. Bracci et al., 1989; DiPasquale and
Cakmak, 1990).
 Use of synthetic ground motions could be explored, where the inherent variability in
ground motion records could be reduced and therefore potentially better isolate the effect
of ground motion duration.
79

 Use of different structural typologies designed to different building codes could be


studied, including effects of higher mode response and soil-structure interaction.

3.9 Acknowledgements

The authors would like to acknowledge the support of Oregon State University and
Endowed Kearny Faculty Scholar that supported the first and second authors. The opinions and
conclusions presented in this paper are those of the authors and do not necessarily reflect the
views of the sponsoring organizations.

3.10 References

Abrahamson, N. A. (1988). “Statistical properties of peak ground accelerations recorded by the


SMART 1 array.” Bulletin of the Seismological Society of America, 78(1), 26–41.

Ancheta, T. D., Darragh, R. B., Stewart, J. P., Seyhan, E., Silva, W. J., Chiou, B. S. J., Wooddell,
K. E., Graves, R. W., Kottke, A. R., Boore, D. M., Kishida, T., and Donahue, J. L. (2014).
“NGA-West2 database.” Earthquake Spectra, 30(3), 989–1005.

Ang, A. H.-S., Kim, W. J., and Kim, S. B. (1993). “Damage Estimation of Existing Bridge
Structures.” Structural Engineering in Naturakl Hazards Mitigation:ASCE Structures
Congress, Irvine, CA, 1137–1142.

ASCE. (2000). FEMA 356 Prestandard: Prestandard and commentary for the seismic
rehabilitation of buildings. Washington D.C.

Baker, J. W. (2007). “Quantitative classification of near-fault ground motions using wavelet


analysis.” Bulletin of the Seismological Society of America, 97(5), 1486–1501.

Baker, J. W. (2015). “Efficient Analytical Fragility Function Fitting Using Dynamic Structural
Analysis.” 31(1), 579–599.

Baker, J. W., and Cornell, C. A. (2005). “A vector-valued ground motion intensity measure
consisting of spectral acceleration and epsilon.” Earthquake Engineering and Structural
Dynamics, 34(10), 1193–1217.

Banon, H., Irvine, H. M., and Biggs, J. M. (1981). “Seismic damage in reinforced concrete
frames.” Journal of the Structural Division, ASCE, 107(9), 1713–1729.
80

Barbosa, A. R., Ribeiro, F. L. A., and Neves, L. C. (2017). “Influence of earthquake ground-
motion duration on damage estimation: application to steel moment resisting frames.”
Earthquake Engineering and Structural Dynamics.

Bojorquez, E., Iervolino, I., Manfredi, G., and Cosenza, E. (2006). “Influence of ground motion
duration on degrading SDOF systems.” Proceedings of 1st European Conference on
Earthquake Engineering and Seismology, Geneve, Switzerland.

Bommer, J. J., Magenes, G., Hancock, J., and Penazzo, P. (2004). “The Influence of Strong-
Motion Duration on the Seismic Response of Masonry Structures.” Bulletin of Earthquake
Engineering, 2(1), 1–26.

Boore, D. M. (2005). “On pads and filters: Processing strong-motion data.” Bulletin of the
Seismological Society of America, 95(2), 745–750.

Boore, D. M., and Bommer, J. J. (2005). “Processing of strong-motion accelerograms: Needs,


options and consequences.” Soil Dynamics and Earthquake Engineering, 25(2), 93–115.

Bozorgnia, Y., and Tso, W. K. (1984). “Inelastic earthquake response of asymmetric structures.”
ASCE Journal of Structural Engineering, 112(2), 383–400.

Bracci, J. M., Reinhorn, A. M., Mander, J. B., and Kunnath, S. K. (1989). Deterministic model
for seismic damage evaluation of RC Structures. Technical Report NCEER-89-0033.

Chai, Y. H. (2005). “Incorporating low-cycle fatigue model into duration-dependent inelastic


design spectra.” Earthquake Engineering and Structural Dynamics, 34(1), 83–96.

Chakroborty, S., and Roy, R. (2013). “Role of Ground Motion Characteristics on Inelastic
Seismic Response of Irregular Structures.” 22(1), 1–16.

Chandramohan, R., Baker, J. W., and Deierlein, G. G. (2016). “Quantifying the Influence of
Ground Motion Duration on Structural Collapse Capacity Using Spectrally Equivalent
Records.” Earthquake Spectra, 32(2), 927–950.

Chester, D. K. (2001). “The 1755 Lisbon Earthquake.” Progress in Physical Geography, 25(3),
363–383.

Chung, Y. S., Meyer, C., and Shinozuka, M. (1987). Seismic damage assessment of RC
members. Technical Report NCEER-87-0022.

Cohee, B. P., Somerville, P. G., and Abrahamson, N. A. (1991). “Simulated ground motions for
hypothesized Mw= 8 subduction earthquakes in Washington and Oregon.” Bulletin of the
Seismological Society of America, 81(1), 28–56.
81

Cornell, A. C. (1997). “Does duration really matter?” Proceedings of the FHWA/NCEER


Workshop on the National Representation of Seismic Ground Motion for New and Existing
Highway Facilities, Burlingame, Calif., organized by NCEER (project 106-F-5.4.1) and
ATC (project ATC-18-1), 125–133.

COSMOS. (2012). “Strong Motion Virtual Data Center (VDC).”


<http://www.strongmotioncenter.org/vdc/scripts/default.plx>.

Darwin, D., and Nmai, C. (1986). “Energy dissipation in RC beams under cyclic load.” ASCE
Journal of Structural Engineering, 112(8), 1829–1846.

De-la-Colina, J. (2003). “Assessment of design recommendations for torsionally unbalanced


multistory buildings.” Earthquake Spectra, 19(1), 47–66.

DiPasquale, E., and Cakmak, A. S. (1990). “Relation between Global Damage Indices and Local
Stiffness Degradation.” ASCE Journal of Structural Engineering.

Dutta, A., and Mander, J. B. (2001). “Energy Based Methodology for Ductile Design pf
Concrete Columns.” ASCE Journal of Structural Engineering, 127(12), 1374–1381.

Dutta, S. C., and Das, P. K. (2002). “Inelastic seismic response of code-designed reinforced
concrete asymmetric buildings with strength degradation.” Engineering Structures, 24(10),
1295–1314.

EN 1992-1-1. (2004). “Eurocode 2: Design od Concrete structures - Part 1.1: General rules and
rules for buildings.”

EN 1998-1. (2004). “Eurocode 8: Design of structures for earthquake resistance—Part 1: General


rules, seismic actions and rules for buildings.” European Committee for Normalization,
Brussels.

Faggella, M., Barbosa, A. R., Conte, J. P., Spacone, E., and Restrepo, J. I. (2013). “Probabilistic
seismic response analysis of a 3-D reinforced concrete building.” Structural Safety, 44, 11–
27.

Fajfar, P. (1992). “Equivalent ductility factors, taking into account low-cycle fatigue.”
Earthquake Engineering & Structural Dynamics, 21(10), 837–848.

Fardis, M., and Negro, P. (2005). “Seismic Performance Assessment and Rehabilitation of
Existing Buildings.” Proceedings of the International Workshop SPEAR, Ispra, Italy.

Filippou, F. C., Popov, E. P., and Bertero, V. V. (1983). Effects of Bond Deterioration on
Hysteretic Behaviour of Reinforced Concrete Joints. Report to the National Science
Foundation. Earthquake Engineering Research Center, Berkeley, California.
82

Foschaar, J. C., Baker, J. W., and Deierlein, G. G. (2012). “Preliminary Assessment of Ground
Motion Duration Effects on Structural Collapse.” Proceedings of 15th World Conference on
Earthquake Engineering, Lisbon, Portugal.

Gosain, N. K., Moore, W. P., and Jirsa, J. O. (1977). “Shear Requirements for Load Reversals on
RC Members.” Journal of the Structural Division, ASCE, 103(7), 1461–1476.

Haddadi, H., Shakal, A., Huang, M., and Parrish, J. (2008). “Report on Progress at the Center for
Engineering Strong Motion Data ( CESMD ).”

Hancock, J., and Bommer, J. J. (2006). “A State-of-Knowledge Review of the Influence of


Strong-Motion Duration on Structural Damage.” Earthquake Spectra, 22(3), 827–845.

Haselton, C. B., Liel, A. B., and Lange, S. T. (2008). “Beam-Column Element Model Calibrated
for Predicting Flexural Response Leading to Global Collapse of RC Frame Buildings.” Peer
2007, 03(May).

Iervolino, I., Manfredi, G., and Cosenza, E. (2006). “Ground motion duration effects on
nonlinear seismic response.” Earthquake Engineering and Structural Dynamics, 35(1), 21–
38.

Jeong, G. D., and Iwan, W. D. (1988). “The effect of earthquake duration on the damage of
structures.” Earthquake Engineering & Structural Dynamics, 16(8), 1201–1211.

Karsan, I. D., and Jirsa, J. O. (1969). “Behavior of concrete under compressive loading.” Journal
of Structural Division, ASCE, 95(ST12), 2543–2563.

Krätzig, W. B., Meyer, I. F., and Meskouris, K. (1989). “Damage Evolution in Reinforced
Concrete Members Under Cyclic Loading.” 5th International Conference on Structural
Safety and reliability (ICOSSAR 89) - Vol II, San Francisco CA, 795–802.

Long, Y. (2012). “Effect of subduction zone earthquakes on SDOF bridge models. MsD Thesis.”
Oregon State University.

Lybas, J. ., and Sozen, M. (1977). Effect of beam strength ratio on dynamic behaviour of
reinforced concrete coupled walls. Report SRS No. 444.

Malley, J. O., Deierlein, G., Krawinkler, H., Maffei, J. R., Pourzanjani, M., Wallace, J., and
Heintz, J. (2010). Modeling and acceptance criteria for seismic design and analysis of tall
buildings. Technical Report PEER/ATC 72-1.

Manfredi, G., Polese, M., and Cosenza, E. (2003). “Cumulative demand of the earthquake
ground motions in the near source.” Earthquake Engineering and Structural Dynamics,
32(12), 1853–1865.
83

Mccabe, S. L., and Hall, W. J. (1989). “Assessment od Seismic Structural Damage.” ASCE
Journal of Structural Engineering, 115(9), 2166–2183.

McKenna, F., Scott, M. H., and Fenves, G. L. (2010). “Nonlinear Finite-Element Analysis
Software Architecture Using Object Composition.” Journal of Computing in Civil
Engineering, 24(1), 95–107.

Menegoto, M., and Pinto, P. E. (1973). “Menegoto-Method of analysis for cyclically loaded R.C.
plane frames including changes in Geometry and non-elastic behavior of elements under
combined normal force and bending.” Proceedings of the symposium resitance and
Ultimate Deformability of Structures Acted on by Well Repeated Loads, Lisboa.

Nassar, A. ., and Krawinkler, H. (1991). Seismic demands for SDOF and MDOF systems. No.
95. John A. Blume Earthquake Engineering Center, Department of Civil Engineering. John
A. Blume Earthquake Engineering Center, Department of Civil Engineering, Stanford
University.

National Research Institute for Earth Science ans Disaster. (1996). “Strong-motion Seismograph
Networks (K-NET, KiK-net).” <http://www.kyoshin.bosai.go.jp/>.

Negro, P., Mola, E., Molina, F. J., and Magonette, G. E. (2004). “Full-scale PSD testing of a
torsionally unbalanced three-storey non-seismic RC frame.” 13th World Conf. Earthq. Eng.,
(968), 1–15.

Oyarzo-vera, C., and Chouw, N. (2008). “Effect of earthquake duration and sequences of ground
motions on structural responses.” Proceedings of the 10th International Symposium on
Structural Engineering of Young Experts.

Park, Y., and Ang, A. (1985). “Mechanistic Seismic Damage Model for Reinforced Concrete.”
ASCE Journal of Structural Engineering, 111(4), 722–739.

Park, Y. J., Ang, a. H. S., and Wen, Y. K. (1987). “Damage-Limiting Aseismic Design of
Buildings.” Earthquake Spectra.

Paulay, T. (1997). “A Review of Code Provisions for Torsional Seismic Effects in Buildings.”
Bulletin of the New Zeland National Society for Earthquake Engineering, 30(3).

Pauley, T., and Priestley, M. J. N. (1992). Seismic Design of Reinforced Concrete and Masonry
Buildings. Wiley.

Popovics, S. (1973). “A numerical approach to the complete stress-strain curve of concrete.”


Cement and Concrete Research, Pergamon, 3(5), 583–599.
84

Powell, G. H., and Allahabadi, R. (1988). “Seismic damage prediction by deterministic methods:
Concepts and procedures.” Earthquake Engineering & Structural Dynamics, 16(5), 719–
734.

Priestley, M. J. N., Calvi, G. M., and Kowalsky, M. J. (2007). Displacement-Based Seismic


Design of Structures. (I. Press, ed.), Pavia.

Raghunandan, M., and Liel, A. B. (2013). “Effect of ground motion duration on earthquake-
induced structural collapse.” Structural Safety, Elsevier, 41, 119–133.

Raghunandan, M., Liel, A. B., and Luco, N. (2015). “Collapse Risk of Buildings in the Pacific
Northwest Region due to Subduction Earthquakes.” Earthquake Spectra, 31(4), 2087–2115.

Ribeiro, F. L. A., Barbosa, A. R., Asce, M., and Neves, L. C. (2014). “Application of Reliability-
Based Robustness Assessment of Steel Moment Resisting Frame Structures under Post-
Mainshock Cascading Events.” 140(2012), 1–12.

Roufaiel, M. S. L., and Meyer, C. (1987). “Analytical Modeling of Hysteretic Behavior of R/C
Frames.” ASCE Journal of Structural Engineering, 113(3), 429–444.

Ruiz-García, J. (2010). “On the influence of strong-ground motion duration on residual


displacement demands.” Earthquake and Structures, 1(4), 327–344.

Shome, N., Cornell, C. A., and Bazurro, P. (1998). “Earthquakes, records, and nonlinear
responses.” Earthquake Spectra, 14(3), 469–500.

Sommer, A., and Bachmann, H. (2005). “Seismic behavior of asymmetric RC wall buildings:
Principles and new deformation-based design method.” Earthquake Engineering and
Structural Dynamics, 34(2), 101–124.

Song, R., Li, Y., and van de Lindt, J. W. (2014). “Impact of earthquake ground motion
characteristics on collapse risk of post-mainshock buildings considering aftershocks.”
Engineering Structures, Elsevier Ltd, 81, 349–361.

Stathopoulos, K. G., and Anagnostopoulos, S. a. (2005). “Inelastic torsion of multistorey


buildings under earthquake excitations.” Earthquake Engineering and Structural Dynamics,
34(12), 1449–1465.

Stephens, J. E., and Yao, J. T. P. (1987). “Damage assessment using response measurements.”
ASCE Journal of Structural Engineering, 113(4), 787–801.

Stone, W. C., and Taylor, A. W. (1993). Seismic performance of circular bridge columns
designed in accordance with AASHTO/CALTRANS standards. National Institute of
Standards and Technology.
85

Tena-colunga, A. (2004). “Evaluation of the seismic response of slender , setback RC moment-


resisting frame buildings designed according to the seismic guidelines of a modern seismic
code.” Proceedings of the 13th World Conference on Earthquake Engineering, Vancouver,
Canada.

Tremblay, R., and Atkinson, G. M. (2001). “Comparative Study of the Inelastic Seismic Demand
of Eastern and Western Canadian Sites.” Eartquake Spectra, 17(2), 333–358.

Valles, R. E., Reinhorn, A. M., Kunnath, S. K., Li, C., and Madan, A. (1996). IDARC 2D
Version 4.0: A Program for the Inelastic Damage Analysis of Buildings. Technical Report
NCEER-96-0010. Technical Report NCEER-96-0010.

Veletsos, A., and Newmark, N. M. (1960). “Effect of inelastic behavior on the response of
simple systems to earthquake motions.” Proceedings of the 2nd World Conference on
Earthquake Engineering.
86

4 DAMAGE-BASED FRAGILITY ASSESSMENT OF RC BRIDGES RETROFITTED


WITH TITANIUM ALLOY BARS ACCOUNTING FOR THE EFFECTS OF
GROUND MOTION DURATION

Andre Belejo, Andre Ramos Barbosa, and Christopher Higgins

Journal of Bridge Engineering


American Society of Civil Engineers
1801 Alexander Bell Drive, Reston, VA 20191-4400
Prepared for Submission
87

4.1 Abstract

The effects of the seismic hazards only started to be considered in bridge design codes
during the last few decades. Therefore, most of the existing vintage bridges were not adequately
designed for the expected seismic hazard when located in seismic zones. In reinforced concrete
(RC) bridge columns, common deficiencies include insufficient lap-splice located in regions
expected to form plastic hinges, as well as flexural and shear reinforcement that are inadequate to
resist the expected seismic demands. In addition, the role of ground motion duration on the
seismic performance of bridges is unknown, and despite the recent focus on the ground motion
duration by the research community, little work has been focused on bridges.

This paper presents results on the effects of ground motion duration and impact of the
retrofit solution employed on the damage-assessment of vintage bridges. Bridges considered
include typical overpasses built in the 1950s – 1970s in the western United States. The models
analyzed include those of as-built bridges and bridges retrofitted with titanium alloy bars (TiAB).
The damage assessment is performed using nonlinear finite element models developed in
OpenSees, in which the finite element components are validated for the retrofitted and non-
retrofitted columns that were part of a testing campaign, which is summarized herein. The
demand and damage parameters considered include peak drift ratios and two damage indices
available in the literature through incremental dynamic analyses and fragility curves. The results
indicated that: (1) the retrofit solution employed to the columns provides a great increase on the
bridges structural capacity, (2) the ground motion duration plays a role in the damage predicted
by the two damage indices in both non-retrofitted and retrofitted bridges, and (3) the retrofit
solution is not as efficient in irregular bridges as when applied to regular bridges.

4.2 Introduction

The impact of severe earthquake conditions on reinforced concrete bridges is typically


governed by the performance of reinforced concrete (RC) columns or the bridge abutments. In
older existing bridges built in the 1950s and 1960s, bridges in most parts of the world, including
the United States of America (USA), were non-seismically designed and often included lap-
splices at the connection between the base of column piers and the foundations. For example,
88

based on the analysis of the Oregon Department of Transportation (DOT) bridge database with
more than 7,000 existing bridges in the State of Oregon, USA, the most common seismic
deficiencies observed for bridges built in the period ranging from the 1950s to the 1970s were
that: (1) lap-splices at the base of the columns had insufficient lengths, and (2) shear detailing
was insufficient and could lead to brittle failure when these are subjected to ground motions.
Several publications addressed the fact that the influence of short lap-splice lengths and
inadequate transverse reinforcement are the main causes of poor performance of vintage bridge
columns subjected to cyclic loads (Cairns and Arthur, 1979; Paulay, 1982; Lukose et al., 1982;
Girard and Bastien, 2002; Melek and Wallace, 2004; ElGawady et al., 2010). To date there are
still many bridges around the world that include the two types of deficiencies mentioned above.
In the State of Oregon, for example, the most common RC bridges that were constructed in the
1950s and 1960s are predominantly reinforced concrete bridges with three to five spans. This
region of the Pacific Northwest of the USA can be subjected to an impending M9.0 earthquake
that may generate strong ground motion with very long durations. Thus, retrofit strategies and
assessment methods for evaluating alternative retrofit strategies are needed.

When dealing with bridge columns that contain lap splices with insufficient length, special
attention should be taken when modeling short lap-splices and their impact on structural
response. Several studies suggested modeling approaches that capture the effect bond-slip
developed in these short lap-splices (Harajli et al., 2004; Zhao and Sritharan, 2007; Harajli,
2009; Chowdhury and Orakcal, 2012). Harajli et al. (2004) proposed a local bond stress–slip
monotonic relationship to characterize the behavior of steel bars when the splitting mode of bond
failure occurs, being later enhanced to account for the cyclic response (Harajli 2009) based on
experimental tests performed. The Harajli (2009) model considers the level of confinement of the
concrete and considers the effect of critical bond parameters such as the diameter of steel bars,
the ratio of concrete cover to bar diameter, the concrete compressive strength, the type of
confinement (steel ties, fiber-reinforced concrete, or fiber-reinforced polymer jackets) and the
area or contact of confining reinforcement. Zhao and Sritharan (2007) introduced a hysteretic
model to capture the slip and strain penetration effects. A reinforcing bar stress versus slip
response relationship was developed to be integrated into fiber-based analyses of concrete
structures using a zero-length section element. The model was validated through comparison of
89

results obtained when applying the model, with measured global and local responses of two
concrete columns and a bridge tee-joint system. Chowdhury and Orakcal (2012) presented a
modeling approach to apply in reinforced concrete columns subjected to horizontal cyclic load,
where a bond stress versus slip relationship was implemented in the formulation of a fiber-based
flexural macro model. This model directly reflected the influence of local bond slip
deformations, due to the local coupling between flexural deformations and slip deformations
associated with either splitting or pullout bond failures along the lap-splice region.

The second topic of interest in this study is related to the probabilistic seismic response and
damage assessment, which can be incorporated into a set of fragility curves. Gidaris et al. (2016)
provided a review of state of the art methods for fragility assessment of bridges. Mackie and
Stojadinović (2005, 2006) developed bridge fragility functions for highway overpass bridges.
Nielson and DesRoches (2007) developed a methodology to generate analytical fragility curves
for highway bridges, showing that the bridge as a system is more fragile than individual
components. In terms of seismic behavior of irregular bridges, Kappos et al. (2002), studied the
effect of pier stiffness on the seismic response of RC bridges with irregular configurations and
concluded that small variations in the pier stiffness may lead to significant changes in the
predominant modes of vibration on irregular bridges. Tamanani et al. (2014) investigated
different ways to balance damage in the bridge to obtain a simultaneous or near-simultaneous
failure of all unequal piers independently of their height, by having suitable arrangements of
transverse reinforcement, while the shortest piers have to be designed for higher ductility
capacity. In terms of retrofitted bridges, Kim and Shinozouka (2004) presented a methodology
for developing and comparing the improvement in the fragility curves of bridges before and after
retrofit. To formulate the problem of fragility enhancement, the quantification was made by
comparing the median values of the fragility curves before and after the retrofit. Padgett and
DesRoches (2009) presented a methodology for developing fragility curves for retrofitted bridge
systems, where importance of evaluating the impact of the retrofit on the overall bridge fragility
is emphasized. The authors presented results on the influence of the different retrofit measures on
the fragility of each class of typical bridges in the Central and Southeastern United States.
90

Even though it could be expected that ground motion duration should play a role in the
seismic assessment of existing and retrofitted bridges, the literature on the topic reports mixed
results. Among several studies over the past decades on the effects of ground motion duration on
structural response (Nassar and Krawinkler 1991; Cornell 1997; Shome et al. 1998; Tremblay
and Atkinson 2001; Dutta and Mander 2001; Manfredi et al. 2003; Bommer et al. 2004; Chai
2005; Bojorquez et al. 2006; Hancock and Bommer 2006; Oyarzo-vera and Chouw 2008; Ruiz-
García 2010; Raghunandan and Liel 2013; Song et al. 2014; Raghunandan et al. 2015;
Chandramohan et al. 2016; Barbosa et al. 2017) only a few studies identified that ground motion
duration had an influence on peak responses, including: (1) higher peak deformations
(Chandramohan et al. 2016; Barbosa et al. 2017), (2) more damage accumulation (Dutta and
Mander 2001; Manfredi et al. 2003; Oyarzo-vera and Chouw 2008; Barbosa et al. 2017; Belejo
et al. 2017); (3) greater residual deformations by Ruiz-Garcia (2010), and (4) increased
probability of reaching collapse (Raghunandan and Liel 2013; Song et al. 2014; Raghunandan et
al. 2015; Chandramohan et al. 2016). On the other hand, in geotechnical earthquake engineering,
the ground shaking duration effects are accepted amongst the research community, where
increased ground motion duration leads to increased soil deformations, increased residual
displacements, and increased probability of liquefaction triggering (Kayen and Mitchell 1997;
Kramer and Mitchell 2006; Sideras and Kramer 2012). The effect of duration in bridge structures
has been addressed in a few recent studies. Chandramohan et al. (2016) studied the effect of
ground motion duration on a RC bridge pier model, in which a reduction in the median collapse
capacity estimated by the long duration set was obtained when compared to the short duration
set. Dusicka and Lopez (2016) developed a model to obtain the acceleration values expected
from a full rupture Cascadia Subduction Zone event and Dusicka and Bazaez (2016a) developed
quasi-static loading protocols that reflect greater inelastic demands of bridge columns subjected
to long duration ground motions caused by subduction megathrust earthquakes. In a companion
paper (Bazaez and Dusicka 2016b) these protocols were applied together with large scale
experiments to evaluate a bridge pier retrofit solution using buckling restrained braces (BRBs).
In addition, other retrofitting solutions have been applied to deficient RC bridge columns during
the last decades: Carbon Fiber-Reinforced Polymers (CFRP) and Glass Fiber-Reinforced
Polymers (GFRP), (McLean and Bernards, 1992; Katsumata et al., 1998; Feng et al., 1999;
91

Iacobucci et al., 2003; Schlick and Breña, 2004; Chang et al., 2004; Endeshaw et al., 2008;
ElGawady et al., 2010), Steel Jackets (Abedi et al. 2010), and more recently Titanium Alloy bars
(TiABs) by Lostra et al. (2017) but none focused on the impact of the ground motion duration on
the response of the bridge structures.

The main objectives of this work are to: (1) understand the impact of the TiABs as retrofit
solution in the seismic behavior of vintage bridges; (2) perform an assessment of the ground
motion duration effects on vintage regular bridges with and without retrofit; (3) gain insight on
how the irregularity of bridges can affect findings on the effects of ground shaking duration and
the performance of the retrofit applied to the columns.
92

Figure 4.1 – Columns tests reported in Lostra (2016): (a) Column C1 (reference column), (b)
Column C2, (c) Column C3, (d) Column C2NL, (e) typical cross-section for column C1, and (f)
typical cross section for columns C2 to C2NL in retrofitted region.
93

4.3 Methodology

Ten different bridge structures that include five different geometries (two regular and three
irregular) are here studied. One non-retrofitted and one retrofitted bridge are considered for the
same geometry. For each case a three-dimensional nonlinear finite element model is developed
in OpenSees. In the non-retrofitted bridges, focus is placed on the development of a
phenomenological model that accounts for the effects of bond-slip of the lap-splice region. In the
retrofitted bridges, the retrofit strategy consists in titanium alloy bars (TiABs), and therefore
focus is placed on the nonlinear modeling of the TiABs. Both non-retrofitted and retrofitted
column modeling approaches are validated through a correlation study between laboratory tests
performed at Oregon State University (Lostra 2016) and the numerical models using OpenSees.

For each bridge model, a database of 64 ground motions is used to assess the effects of
ground motion duration on the structural behavior. The selection of bi-directional ground motion
records was performed following the approach for 3-D structures proposed by Belejo et al
(2017), where the spectral matching in both orthogonal directions is performed. In total, 32 long-
and 32 short-duration bi-directional ground motion records with similar 5 percent damped linear
response spectra within the period ranges of interest (0.2T1 to 3T1, where T1 is the fundamental
period of the structure) are selected. In this work 3-D structural models of bridges are subjected
to orthogonal horizontal short- and long-duration ground motion records. An earthquake is
classified “long” when at least one of its orthogonal components significant duration (D5-75) is
greater than 25 seconds (Chandramohan et al. 2016). The ground motions databases used and the
methodology proposed by Belejo et al (2017) are used to isolate the effects of duration. In this
approach, for each long duration ground motion, a short-duration ground motion with the scale
factor SF and rotation angle θj that, together, lead to the smallest difference when compared with
the long-duration ground motion spectral shape, is selected. In this study, it is important to note
that the retrofit applied to the columns has direct impact on the calculation of the ultimate
rotation (Haselton et al. 2008; Malley et al. 2010) due to the differences on the level of
confinement of the concrete that leads to a higher compression strength, transverse reinforcement
and longitudinal reinforcement provided by the TiABs.
After calculation of the damage indices, damage index-based fragility curves that indicate
the probability of reaching a determined damage level based on the classification proposed by
94

Ang et al. (1993) are computed. These functions are based on (1) the Park and Ang damage
index (DIP&A), since is the best known and widely used, and (2) the Reinhorn and Valles damage
index (DIR&V), which is an extension of the Park and Ang that considers the effect of low-cycle
fatigue. The formulation for both DIP&A and DIR&V including the methodology followed on their
calculation and the computation of these damage-index based fragility functions are described in
detail in Belejo et al. (2017).

The main difference between short- and long-duration ground motions lies on the number of
repeated loading cycles, that the structures are subjected to, which may lead to greater energy
dissipated by yielding of the structural members. Therefore, the nonlinear finite element models
used in analysis need to be validated and calibrated such that the amount of energy dissipated is
identical to that observed in testing. In such cases, the damage indices such as the ones by Park
and Ang and Reinhorn and Valles are adequate for studying the effects of ground motion
duration.

It is important to note that both damage indices considered in this study include a component
that accounts for the deformation levels achieved during the analyses as well as a parameter that
specifically accounts for amount of energy dissipated. Park et al (Park et al. 1987), during the
calibration process of the DIP&A tested excitations with different duration. For the excitation with
shorter duration, the damage was caused primarily by excessive deformation, whereas for the
longer duration excitation the effect of repeated cyclic loading became dominant.

The DIR&V, based on DIP&A, also has in its formulation a component related with the energy
dissipated by yielding, which was modified to account for low cycle fatigue.

4.4 Case Studies

The cases of studies considered in this study were chosen to represent the most common
characteristics of existing bridges constructed in the period ranging from 1950 to 1970 in the
Pacific Norwest of the United States. Figure 4.2 provides an overview of the geometry and
dimensions adopted for the ten bridge designs analyzed in this study. These include five
prototype bridges that represent three- and four-span as-built bridges with columns containing
typical detailing shown in Figure 4.1. The first bridge model, RB1, simulates the old Mackenzie
95

River Bridge that was replaced recently. The bridge was located on Highway Interstate 5, near
Eugene, Oregon. The RB1 bridge model is a 3-span bridge, with equal 15.2 m (50 ft) long spans.
Each of the two bridge bents consist of two columns with a clear height of 6.71 m (22 ft). The
second bridge is a prototype design, which represents the median bridge geometry of 7,000
bridges from the Oregon Department of Transportation bridge database. The prototype design
corresponds to a 4-span bridge, with each span being 15.2 m (50 ft) long. The column lengths for
bents 1, 2 and 3 correspond to the median length of all columns from the Oregon DOT bridge
database with 4.88 m (16 ft) clear height. Columns with this length are designated hereforth as
the “median column.” The other three bridge designs considered in this study consist of designs
with varying column lengths. The shortest and longest columns (henceforth “short column” and
“long column”) are defined by the 16th and 84th percentiles of the height of all bridge columns
found in the Oregon DOT bridge database. Thus, the clear height of a short column, median
column, and long columns are 3.05 m (10 ft), 4.88 m (16 ft), and 6.71 m (22 ft), respectively. As
shown in Figure 4.2 (f), bridge IB1 shows a “123” format, where each of the digit in the three
digit number corresponds to the column type assigned to each bent from left to right, in which
“1” corresponds to a “short column”, “2” corresponds to a “median column,” and “3” to the
“long column.” It can be seen in this figure that bridges IB2 and IB3 are characterized by 132
and 213 format, respectively.

Apart from the column lengths, all dimensions and detailing of the original Mackenzie River
Bridge were used to define the prototypes of all bridges used in this study, including: (1) the
column sections are 0.61m (2 ft) wide square columns with four #36M (#11) vertical ASTM 305
Gr. 40 reinforcing steel bars with the nominal yield stress being fy equal to 276 kPa (40 ksi) with
a lap-splice length equal to 0.91 m (3 ft), and #10M (#3) hoops spaced at 0.30 m (12 in) on
center; (2) the section of the girders of the superstructure are 1.22 x 0.37 m (4 ft x 1ft-2.5 in).
with #36M (#1) longitudinal bars and #13M (#4) stirrups as shown in Figure 4.2 (d) with ASTM
305 Gr. 40 reinforcing steel bars. The cap beams are 1.91 m (6 ft – 3in.) high and 0.42 m (1 ft –
4.5 in.) wide, are reinforced with #36M (#11) longitudinal bars and #16M (#5) stirrups. The cap
beams over the abutment bents are 1.27 m x 0.30m (4 ft – 2 in. x 1 ft) reinforced with #25M (#8)
longitudinal bars and #13M (#4) stirrups, as shown in Figure 4.2 (e). In all bridges, the abutment
96

bents are composed by a cap beam supported by four buried columns, therefore in this study the
bridge deck is assumed to be pinned as described in detail in section 4.5.4.

Each of the five bridge types was used to generate models of bridges retrofitted with the
solution proposed in Lostra (2016), which makes use of TiAB longitudinal bars. In all retrofitted
bridge models considered, the length of the TiAB longitudinal bars is 1.52 m (5 ft), measured
from the top face of the footing.
97

(a)

(b)

(d)

(c) (e)
Bridge H1 (m) H11 (m) H2 (m) H21 (m) H3 (m) H31 (m)
RB1 6.65 - 6.65 - - -
RB1R 6.65 1.52 6.65 1.52 - -
RB2 4.85 - 4.85 - 4.85 -
RB2R 4.85 1.52 4.85 1.52 4.85 1.52
IB1 3.00 - 4.85 - 6.65 -
IB1R 3.00 1.52 4.85 1.52 6.65 1.52
IB2 3.00 - 6.65 - 4.85 -
IB2R 3.00 1.52 6.65 1.52 4.85 1.52
IB3 4.85 - 3.00 - 6.65 -
IB3R 4.85 1.52 3.00 1.52 6.65 1.52

(f)

Figure 4.2 - Bridge geometry considered for the study cases: (a) Bridge RB1/RB1R
elevation view; (b) Bridges RB2/RB2R, IB1/IB1R, IB2/IB2R and IB3/IB3R elevation view; (c)
Bent 2 - Section DD’; (d) Deck reinforcement; (e) Cap beam reinforcement, (f) Column and
retrofit lengths assumed for all columns.
98

4.5 Model Development and Validation

4.5.1 Column Modeling Considering a Modified Lap-splice Bond Stress-slip Model


for Use in Fiber Sections
A bond stress-slip model for the lap-splice response is proposed to be used as part of the
cross-sections of displacement-based distributed plasticity finite elements implemented in
OpenSees. The developed bond stress-slip model is integrated at the end of the element along the
length of the lap-splice. The material model for the bond-slip relationship is applied to the fibers
using the Pinching4 model (Lowes et al. 2003) available in OpenSees.

4.5.1.1 Bond stress-slip model parameter calibration


The methodology proposed to develop the stress-strain backbone curve of the lap-splice
component consists in the following steps: (1) adopt a bond stress–slip relationship, (2) convert
the bond stress–slip relationship to a force–deformation constitutive model, (3) assign the force-
deformation constitutive model to a zero-length (spring) element, and connect the spring element
with a truss element with the length of the lap-splice, and (4) perform a monotonic displacement
controlled nonlinear analysis along the axis of the lap-splice model.

Step 1: The bond stress–slip relationship used here follows the model proposed by Harajli et
al. (2004) represented in Figure 4.3 (a). The backbone curve in Harajli et al. (2004) model is
defined by:
0.3
s
u  um   (4.1)
 s1 

where um  2.57 f c' (4.2)

with um equal to the maximum bond stress that corresponds to pullout failure and s1 the
maximum slip mobilized by unconfined concrete that is defined as 15% of the clear distance
between the ribs of the reinforcing bar. The model follows the relationship defined in:
2/ 3
 c  Kc 
usp  0.78 f 
c
'
  um (4.3)
 db 
99

u p  usp  0.5  Kcs  (4.4)

u 
ssp  s1e3.3lnus um   0.4 ln  m 
u 
 sp  (4.5)

where usp is the splitting bond stress and ssp is level of slip corresponding to the splitting bond
stress. The parameters Kc and Kcs in the calculation of the splitting bond stress usp, and the post-
splitting strength up are confinement parameters that depend on the number of spliced tension
bars in the section, cover, spacing between bars and transverse steel area, and the parameter c is
the smallest of the side cover, bottom cover, or half the clear spacing between the bars.
Expressions for Kc and Kcs are given in Harajli et al. (2004).

According to Harajli et al. (2004), the bond-slip model follows a equation (4.1) until the
initial splitting bond stress uspi is reached, which corresponds to 0.7usp., then followed by a linear
segment until the usp. Beyond the ssp level of slip corresponding to usp, the bond stress drops to
the post-splitting strength up, defined in equation (4.4), and then keeps decaying until the
maximum slip smax is reached, which corresponds to a level of slip at which the bond stress
reduces to zero. In this model, smax is equal to the total clear distance between the ribs of the
reinforcing steel bars.

Step 2: To convert the bond stress–slip relationship to a force–slip relation, the bond force
per unit length fB that can be transferred without the assistance of special transverse
reinforcement confining the splice is first computed. This unit force fB is given by

fB  p  u (4.6)

where u is the bond stress and p is the perimeter of the characteristic block defined by the rupture
path observed when the splitting rupture mode occurs.
The peak force in each bar is given by:
ls
2
Fbar   f B dx  umax  p  ls (4.7)
0
3
100

where as shown in Figure 4.3 (d), the shear stress distribution is assumed to follow a semi
parabola from the maximum stress towards zero stress condition at the reinforcing steel bar tip.
Illustratively, Figure 4.3 (b) presents an example of observed splitting failure during a test, which
is schematized in Figure 4.3 (c) to identify the perimeter of the characteristic block p.
Step 3: A series model is developed to capture the combined effects of the bond-slip and the
reinforcing steel bar strength and stiffness. The model consists in: (1) a zero-length element
where the bond force-slip spring is assigned, and (2) a truss element where the reinforcing steel
bar properties are assigned. The Menegotto-Pinto model (Menegoto and Pinto 1973) was
assigned to the truss element, and the area of the truss element was set to equal the effective
reinforcing steel area.
Step 4: A displacement controlled nonlinear static analysis is performed by applying a unit
pulling force to the free node of the truss element along the axis of the bar. Finally, the modified
lap-splice bond stress and strain model is obtained by dividing the forces and displacements
obtained from the nonlinear static analysis by the area of the bars and lap-splice length,
respectively.
101

(a)

(c)

Fbar 2
 umax l s
p 3

l s

(b) (d)

Figure 4.3 - (a) Bond stress – slip monotonic relationship (Harajli et al. 2004); (b) Splitting
of concrete during a laboratory test (Lostra 2016), as an example of the perimeter of the
characteristic block typically considered for the splitting failure mechanism; (c) Scheme for the
definition of the perimeter of the characteristic block considered for the splitting failure
mechanism; (d) Assumed stress distribution along steel bars
102

4.5.1.2 Calibration of the hysteretic parameters for the bond stress-slip model
To define the cyclic behavior for the lap-splice, the model proposed by Harajli (2009) was
adopted. Four additional parameters need to be defined, including: (1) the ratio of the
deformation at which reloading begins to the peak deformation, set to 0.48; (2) the ratio of the
force at which reloading begins to the force on the backbone curve corresponding to peak
deformation achieved on a previous cycle that is set to 1.0; (3) the ratio of strength developed
upon unloading from negative load to the minimum strength developed under monotonic loading
that is set to -0.2; and, (4) the cyclic degradation model parameters for unloading and reloading
stiffness degradation that is set to 0.5. The definition of these parameters followed the
recommendations provided by Harajli (2009), except for the third mentioned parameter that was
calibrated to match the column testing results during the validation process described below.

4.5.1.3 Column model


The column model accounting for the modified lap-splice bond stress-strain model was
discretized as a fiber-section displacement-based beam-column element. Non-retrofitted columns
were divided into five (5) elements: one element along the lap-splice length and four elements
above the lap-splice. For all elements, the cross-section was discretized using 144 (12 x 12)
fibers. Two different concrete models were used. The model used to capture the response of the
confined concrete follows a compressive stress-strain envelope that is based on the Fujii concrete
model (Hoshikuma et al. 1997). This model in OpenSees is known as “Concrete with beta,” and
is a uniaxial concrete material that considers the effect of the strain normal to the axis of the
element in which the material is used. This material model has two options regarding the strength
degradation in tension: (a) tri-linear model or (b) nonlinear model based on the tension stiffening
relation of Stevens et al. (1991). Here the nonlinear tension stiffening model was used. The
softening behavior in compression is tri-linear. A second model based on the Kent-Scott-Park
(Scott et al. 1982) model with modifications proposed in Yassin (1994) was assigned to the
unconfined concrete. This model in OpenSees is known as Concrete02. While this model is
useful and has been shown to be robust, it does not allow for explicit definition of the initial
tangent stiffness. The Menegotto-Pinto (Steel02 in OpenSees) material model as proposed in
Filippou et al. (1983) was assigned to the reinforcing steel bars above the lap-splice.
103

4.5.1.4 Validation of column model response


The finite element model of the column fixed in the base was subjected to a displacement
controlled pushover analysis. The history of displacements applied at the column tip are the
measured displacements applied in reverse cyclic tests of the specimens from Lostra (2016).

Column C1 corresponds to a column that was built to mimic the 1950s detailing of bridge
columns including the short lap-splices. A vertical load of 890 kN was applied to the column and
a truss element was added connecting the base to the top of the column to simulate the Dywidag
rods that were used to apply the gravity loads in the Lostra (2016) experiments.

In Figure 4.4 (a), the results obtained from numerical analysis of column C1 are compared
with the results measured during testing. The marked points on the load–displacement plot are
reference points where the energy dissipated by yielding was calculated. Figure 4.4 (b) shows the
comparison between the energy dissipated by yielding during the numerical analysis and during
the experiment. From analysis of Figure 4.4 (a), it can be observed that both experiment and
numerical model show similar results in terms of initial stiffness, peak capacity for both positive
and negative directions of the load, and pattern of capacity degradation. Figure 4.4 (b) shows
good correlation between the evolution of the energy dissipated during the analysis and the one
dissipated during testing. This fact is extremely important in the validation of the response since
the dissipated energy plays an important role when performing assessments of ground motion
duration effects on the structural response as demonstrated by several authors (Dutta and Mander
2001; Manfredi et al. 2003; Oyarzo-vera and Chouw 2008; Barbosa et al. 2017). The numerical
model for columns C1 is thus able to capture the response of non-retrofitted column subjected to
cyclic loads.

4.5.2 TiAB Retrofit Modeling

4.5.2.1 Material characteristics


The titanium alloy used here is the TiAl6V4. It is characterized by having high strength and
low stiffness when compared with conventional reinforcing steel bars, and by its low or
negligible hardening. TiAB has a yield stress of approximately 1,000 MPa (140 ksi), and
Young’s modulus equal to 107 GPa (15,500 ksi).
104

4.5.2.2 Column modeling, calibration, and validation


From the laboratory testing and reported in Lostra (2016), one column without lap-splices
was included, designated here as C2NL. In column C2NL, the longitudinal column reinforcing
steel bars were not anchored into the footing, starting immediately above the footing. The load
transfer in tension was provided by the TiABs only. Therefore, in the model developed in this
subsection, the contribution of the reinforcing steel bars is neglected. However, since the
concrete spiral and concrete shell provide additional confinement, the modeled concrete square
column is assumed to be confined. The confining concrete parameters were calculated following
Mander et al. (1988) guidelines.

The vertical TiABs are modeled as truss elements that are connected to the displacement-
based finite element of the concrete column through rigid links as shown in Figure 4.5 (b). The
Pinching4 model in OpenSees is used to model the hysteretic properties of the TiABs. The
calibration of the hysteretic parameters was performed through trial-and-error to correlate the
column C2NL test results with the numerical counterparts. Figure 4.4 (c) shows the test results
and numerical results obtained from calibrated model, where it can be seen that the model show
good correlation with the test results.

4.5.3 Combined Lap-splice and TiAB Retrofit Modeling


The numerical models developed in the previous sections are here combined to obtain a
complete numerical model for the retrofitted column that includes the splice (Column C2).

4.5.3.1 Validation
Figure 4.4 (e) shows the load-displacement relationships obtained from the numerical
analysis and from the experiment data of column C2. It can be seen that the model captures well
the initial stiffness and the peak capacity. Figure 4.4 (f) shows the evolution of the hysteretic
energy dissipated. It can also be seen here a good correlation between model and test results. In
summary, the numerical model of column C2 seems to capture well the structural behavior of the
retrofitted column under dynamic loads.
105

200 400

Energy dissipated,
Column C1
100 300

kNm.rad
Applied Load, kN

0 200
Experiment data
Experiment
-100 data 100
OpenSees OpenSees
model model
-200 0
-0.3 -0.15 0 0.15 0.3 0 0.1 0.2 0.3
Top displacement, m Top displacement, m

(a) (b)
200 1500
Column C2NL 1 9
1000
Applied Load, kN

100 8 2
Stress, MPa 500
0 0
7

-0.2 -0.1 0 3 0.1 0.2


-100 6
-500
4
5
-1000
-200
-0.4 -0.2 0 0.2 0.4 -1500
Top displacement, m Strain, m/m

(c) (d)
300 1500
Column C2
Energy dissipated,

150
Applied Load, kN

1000
kN.m.rad

-150 500

-300
-0.3 -0.15 0 0.15 0.3 0
Top displacement, m 0 0.1 0.2 0.3
Top displacement, m

(e) (f)

Figure 4.4 – Validation of component response. Results measured in the laboratory


experiments and in the analytical model: (a) Column C1 load-displacement cyclic response; (b)
Column C1 – hysteretic energy dissipated; (c) Column C2NL load-displacement cyclic response;
(d) ) Response of TiABs to one load cycle; (e) Column C2 load-displacement cyclic response ;
(f) Column C2 – hysteretic energy dissipated.
106

4.5.4 Bridge Models


Figure 4.5 (a) shows the main components considered in the bridge models developed
herein, including: (1) the superstructure is modeled using linear elastic beam-column elements,
with its stiffness reduced 50% due to typical cracking that is often observed in these 1950s –
1970s existing bridges, following recommendations of ACI 318 (ACI 2011); (2) the connections
between the columns and cap beams are considered to be monolithic and rigid; (3) the cap beams
are modeled as rigid elements; (4) the columns are fixed at the base, not considering soil-
structure interaction effects in this study; (5) at the abutments, zero-length elements are
connected to the abutment cap beam elements, which provide the passive resistance of the soil in
contact with the cap beam modeled through a HyperbolicGap material (Duncan and Mokwa
2001). The parameters defining the HyperbolicGap model are initial stiffness,
unloading/reloading stiffness, ultimate (maximum) passive resistance, initial gap, and failure
ratio. The failure ratio Rf is defined as the ratio between the ultimate passive pressure load and
the hyperbolic asymptotic value of passive resistance. Rf can be obtained when measuring the
response, or estimated based on experience.

The parameters assigned to the zero-length elements connected to the abutment cap beams
are assigned considering the tributary area of each element. The passive resistance material
model initial stiffness was obtained following Douglas and Davis (1964). The ultimate passive
resistance is calculated based on the Rankine theory. The abutment soil characteristics were
obtained from the report SSRP-05/02 by Earth Mechanics (2005) where field investigation for
abutment backfill characterization was performed. Among the typical backfills, medium dense
silty sand was the one assumed for the abutment backfill. A failure ratio of 0.7 and an initial gap
of 0.025m were assigned to the zero-length elements along the longitudinal direction, and no gap
was assigned to the elements resisting along the transverse direction. The initial gap along the
longitudinal direction and the failure ratio values were defined according to abutment stiffness
models proposed for bridge simulations by Wilson and Elgamal (2006), which are based on
large-scale abutment tests performed on the shaking table at UC San Diego.
107

Figure 4.5 – Conceptual representation (a) Nonlinear finite element model developed in
Opensees; (b) Column model with TiAB retrofit.

4.5.5 Ground Motion Selection


A total of ten bridge models were developed, as shown in Figure 4.2. Since different column
lengths defined, the bridges studied here show different fundamental periods. The set of long-
duration ground motions was the same for all bridges, however when applying the selection
procedure described in Belejo et al. (2017) to select short-duration ground motions, two different
sets were selected. First, one short-duration ground motion set was selected for the study of the
three-span bridges (RB1 and RB1R) in which the fundamental period ranged from 1.25 sec ≤ T1
< 1.30 sec. Second, for the four-span bridges (RB2, RB2R, IB1, IB1R, IB2, IB2R, IB3 and
IB3R), in which the fundamental periods ranged between 0.74 sec and 0.85 sec. The pairs of
long-duration and short-duration earthquake records selected for the three-span bridges and four-
span bridges with information regarding to scale factors and rotation angles are listed in Tables
A4 and A5 respectively.
Tables A6, A7 and A8 show information regarding both components of the long-duration
and scaled short-duration sets of records selected in terms of significant duration D 5-75, rupture
distance, peak ground acceleration (PGA), moment magnitude (MW), and the 5-percent damped
linear elastic spectral acceleration at the fundamental period of the structure Sa(T1). To assure that
spectral accelerations and significant durations are not correlated at any nonlinearity level of the
108

structures, Figure 4.6 (e) and Figure 4.6 (f) show the median correlation between 5-percent
damped linear elastic as a function of period for the three-span and four-span bridges,
respectively. It can be seen that there is little to no correlation, so the duration effect can be
considered an isolated variable in the subsequent analyses.
109

(a) (b)

(c) (d)

(e) (f)

Figure 4.6 - (a) Components #1 of the response spectra of short- and long-duration grounds
motions of pair #6 (Table A4); (b) Components #2 of the response spectra short- and long-
duration grounds motions of pair #6 (Table A4); (c) Geometric Mean of the response spectra of
short- and long-duration grounds motions of pair #6 (Table A4); d) Accelerograms of the
components #1 of pair #6 (Table A4); (e) Correlation coefficient between D5-75, and natural
logarithm of the Sa(T1) versus period of vibration, considering the ground motions selected for
Bridge RB1/RB1R; (f) Correlation coefficient between D5-75, and natural logarithm of Sa(T1)
versus period of vibration considering the ground motions selected for Bridge RB2/RB2R,
IB1/IB1R, IB2/IB2R, and IB3/IB3R
110

4.6 Results

4.6.1 Modal Results


Figure 4.7 (a) and (b) show the deformed shapes for the two first modes of bridges
RB1/RB1R and RB2/RB2R respectively, while mode shapes and periods of bridges IB1/IB1R,
IB2/IB2R and IB3/IB3R are displayed in Figure 4.7 (c) - (e). In all cases, the bridge first mode is
characterized by translation along the longitudinal direction, while in the second mode the bridge
deforms along its transverse direction. The bridge RB1 is the most flexible with its mode shapes
showing pure translation in both directions. Bridge RB2 also shows pure translation in both
mode shapes with higher fundamental period when compared with the irregular bridges. It is
interesting to note that bridge IB3 shows a smaller second period when compared to bridges IB1
and IB2 due to the fact that, in bridge IB3, the short column is located in the middle of the bridge
(position “2”) restrained the modal displacement of the deck due to the greater stiffness of the
short columns. On the other hand, in bridges IB1 and IB2 the short column is located at one of
the bridge ends, restraining the motion of the spans supported by the short column.
111

T1 = 1.26 sec T1 = 0.85 sec


First Mode RB1 First Mode RB2
T1 = 1.25 sec T1 = 0.85 sec
RB1R RB2R

T2 = 1.02 sec T2 = 0.56 sec


Second Mode RB1
Second Mode RB2
T2 = 1.04 sec T2 = 0.56 sec
RB1R RB2R

(a) (b)

First T1 = 0.74 sec First T1 = 0.74 sec First T1 = 0.74 sec


IB1 IB3
IB2
Mode T1 = 0.74 sec Mode T1 = 0.74 sec Mode T1 = 0.74 sec
IB1R IB3R
IB2R

Second T2 = 0.60 sec Second T2 = 0.60 sec Second T2 = 0.53 sec


IB1 IB2 IB3
Mode T2 = 0.60 sec Mode T2 = 0.60 sec Mode T2 = 0.53 sec
IB1R IB2R IB3R

(c) (d) (e)

Figure 4.7 - First mode shapes and respective periods: (a) Bridge RB1/RB1R; (b) Bridge
RB2/RB2R; (c) Bridge IB1/IB1R; (d) Bridge IB2/IB2R; (e) Bridge IB3/IB3R

4.6.2 Time-history Response Analysis

4.6.2.1 Deformation and Energy Demands


The time-history drift ratios of column C1-1 of RB1/RB1R, RB2/RB2R and IB1/IB1R are
computed and shown in Figure 4.9 where the effect of the retrofit can be assessed in terms of
112

peak drift ratios. The effect of the retrofit is also assessed by computing the energy dissipated of
the bridge columns along the time of the ground motion. Plots showing the evolution of energy
dissipated by yielding of columns C1-1 of RB1/RB1R, RB2/RB2R and IB1/IB1R are shown in
Figure 4.9. Figure 4.10 (a) and (b) show how the amount of energy dissipated by the short and
median columns vary when located in different positions along the bridge, without and with
retrofit respectively, while Figure 4.10 (c) and (d) show the difference of energy dissipated by
yielding in the median column during a short- and long-duration ground motion. To better
understand the difference of energy dissipated by the short and long columns, the response in
terms of shear vs drift ratio of the columns is shown in Figure 4.11 comparing columns C1-1 and
C3-1 of bridge IB1.

(a) (b)

(c) (d)

(e) (f)

Figure 4.8 – Time-history of drift ratios of the non-retrofitted -and retrofitted column C1-1:
(a) Bridge RB1/RB1R along the longitudinal direction; (b) Bridge RB1/RB1R along the
transverse direction; (c) Bridge RB2/RB2R along the longitudinal direction; (d) Bridge
RB2/RB2R along the transverse direction; (e) Bridge IB1/IB1R along the longitudinal direction;
and (f) Bridge IB1/IB1R along the transverse direction;
113

(a) (b) (c)

(d) (e) (f)

Figure 4.9 –Energy dissipated at the top column along between non-retrofitted -and
retrofitted column C1-1: (a) Bridge RB1/RB1R along longitudinal direction; (b) Bridge
RB1/RB1R along transverse direction; (c) Bridge RB2/RB2R along longitudinal direction; (d)
Bridge RB2/RB2R along transverse direction; (e) Bridge IB1/IB1R along longitudinal direction;
and (f) Bridge IB1/IB1R along transverse direction;.
114

(a) (b)

(c) (d)

Figure 4.10 - Energy Dissipated in the column top along longitudinal direction: (a) Different
positions of short “1” and median length columns “2” (Bridges IB1, IB2 and IB3) for Sa(T1) =
0.4g; (b) Different positions of short “1” and median length columns “2” (Bridges IB1R, IB2R
and IB3R) for Sa(T1) = 0.8g; (c) Comparison between short and long direction in Bridge RB1; (d)
Comparison between short- and long-durationin Bridges RB2 and IB1.
115

50
Column Shear, kN

Short column Long column


-50
-0.04 0 0.04 -0.04 0 0.04
Drift Ratio Drift Ratio

(a) (b)
Figure 4.11 - Column shear vs drift ratio time-history response along the longitudinal direction
of bridge IB1: (a) column C1-1 (short column); (b) column C3-1 (long column)

The time-histories of the drift ratio due to the the short-duration ground motion of pair 6
acting on these bridges is shown in

Figure 4.8 (a). This figure shows that the retrofit is able to provide to the capacity for
columns to return to the original position, due to self-centering ability provided by the retrofit
solution employed. On

Figure 4.8, the plots of the time-history drift ratios show residual displacements of column
C1-1, which it is not observed for the retrofitted bridges. Figure 4.9 shows that for high seismic
intensities, when the non-retrofitted bridge model shows evidence of high level of nonlinearity,
the retrofitted column shows greater ductility, also leading to greater amount of energy
dissipated.

In Figure 4.10 (a) and (b), the difference of energy dissipated when comparing the short with
the median column is evident. The short columns are naturally subjected to greater forces due to
their larger rigidity and deformation demands. Greater drift ratios are obtained for the short
column when compared with the other columns since the bridge deck moves as a rigid body.
Figure 4.11 shows the hysteretic behavior in terms of shear vs drift ratio for both short and
median columns and the aforementioned differences related with force and deformation demands
between both columns are evident. The results for both columns show that changing position
along the bridge does not affect the column structural behavior in terms of energy demands. The
little differences captured reveal that: (1) the energy dissipated at the short column is a little
116

higher when located at the middle of the bridge, and (2) the energy dissipated at the median
column is slightly higher when positioned two spans away from the short column. These
observations agree with the discussion in section 4.6.1 highlighting that the most important role
is played by the short column in the seismic behavior of the irregular bridges studied. However,
the differences between the irregular bridges IB1/IB1R, IB2/IB2R and IB3/IB3R are minimal
and hence overlooked. Only the results relative to bridge IB1 are shown, being implicit that the
results relative to bridges IB1 and IB1R also reproduce results relative to bridges IB2/IB3 and
IB2R/IB3R respectively.

The differences of energy dissipated when subjected by a short- or long-duration ground


motion are significant. Figure 4.10 (c) and (d) show that the energy dissipated at top of the
median column, when subjected by the long duration ground motion, is much higher that the
energy dissipated by the short-duration ground motion. Results relative to bridge RB1 are
isolated in Figure 4.10 (c) due to the fact that a different set of short-duration ground motions
was considered due to its higher periods as already explained. Figure 4.10 (d) is thus able to
show the impact of bridge irregularity in terms of geometry, when comparing equal length
columns. In bridge RB2 the energy dissipated at any column (all have the same length) is
considerably higher than the energy dissipated by the median column of the irregular bridge with
equal length, explained by the fact that in the irregular bridge, the greater amount of energy
dissipated is concentered in the short column as can be seen in Figure 4.10 (a) and (b).

4.6.2.2 Statistical assessment of the ground motion duration effects


In this section, the damage indicators such as peak drift ratios, Park & Ang damage indices
(DIP&A), and the Reinhorn & Valles damage indices (DIR&V) are estimated using all the unscaled
ground motions considered in this work in order to perform a preliminary assessment of the
ground motion duration effects on the bridges.

, Figure 4.13 and Figure 4.14 show the three damage indicators versus acceleration at the
fundamental period Sa(T1) and significant duration D5-75. In each figure, 128 points are shown
which correspond to the results for each of the 64 x 2 analysis performed. In addition, a multiple
linear regression surface that best fits the data is shown, along with a horizontal line which caps
the limit of the peak response measured.
2 2
2
R = 0.73 R = 0.76 R = 0.81 4

4 3

2 2
1
0
2
100 0 100 2 100 0

Peak Drift Ratio, %


-1 10
50 10 0 10 50 0
-2 10 50
0 0 -2 10 -2
10
10 10 0
10

(a) (b) (c)


2 2 2
R = 0.86 R = 0.45 R = 0.82
4
4
3
2 2
0 1
100 0

Peak Drift ratio, %


50 -1 10 100 2 2
0 100 0
0
0 -2 10 50 10 50 10
0 -2
10 10
10 -2
10 0
10

(d) (e) (f)

Figure 4.12 - Surface plots that relate D5-75, Sa(T1) Peak drift ratio: (a) Bridge RB1; (b) Bridge
RB2 (c) Bridge IB1; (d) Bridge RB1R; (e) Bridge RB2R; (f) Bridge IB1R
117
2 2
2 2
R = 0.77 R = 0.80 R = 0.85 R = 0.82 0.8

0.6
1.0
0.4
0.5

DIP&A
0.0 0.2
100 0 100 2
-1 0 100 2
50 10 50 10 50 0 10 100 2 0.0
0 -2 10 0
0 10-2 10 010-2 10
10 50 -2
10
0 10
10

(a) (b) (c) (d)


2 2 2
2 0.8
R = 0.87 R = 0.85 R = 0.89 R = 0.88
0.4
0.6
0.2

DIP&A
0.4
0.0
100 0 100 0 0.2
50 -1 10 50 -1 2 100 2
0 -2 -2
10 100 0 0
10 10 50 -2 10 50 10
10 0 10
0 10 10 0 10-2 10 0.0

(e) (f) (g) (h)

Figure 4.13 - Surface plots that relate D5-75, Sa(T1) Damage Index Park and Ang: (a) Bridge
RB1; (b) Bridge RB2 (c) Bridge IB1; (d) Most damaged element of Bridge IB1; (e) Bridge RB1R;
(f) Bridge RB2R (g) Bridge IB1R; (h) Most damaged element of Bridge IB1R
118
2 2 2 2
R = 0.66 R = 0.72 R = 0.80 R = 0.79
0.8
1.0 0.6
0.5 0.4

DIR&V
0.0 0.2
100 0 100 2
0 100 2
50 -1 10 50 10
-2 10 50 0 10 100 2 0.0
0 10 0 -2
10 10 0 -2 10 50 0 10
10 0 -2 10
10

(a) (b) (c) (d)


2 2 2 2 0.8
R = 0.79 R = 0.82 R = 0.86 R = 0.86
0.6
0.4
0.4
0.2

DIR&V
0.2
0.0
0
2 0.0
100 50
-1 10 100 100 2
0 -2 50 0 10 100 2
10 10 -2 50 0 10
0 10 50 0 10
10 10
0 10-2
0 -2 10
10

(e) (f) (g) (h)

Figure 4.14 - Surface plots that relate D5-75, Sa(T1) and Damage Index Reinhorn and Valles: (a)
Bridge RB1; (b) Bridge RB2 (c) Bridge IB1; (d) Most damaged element of Bridge IB1; (e) Bridge RB1R;
(f) Bridge RB2R (g) Bridge IB1R; (h) Most damaged element of Bridge IB1R
119
120

shows six plots with peak drift ratios obtained for the unscaled analyses in the bridges
studied: (a) RB1, (b) RB2, (c) IB1, (d) RB1R, (e) RB2R, and (f) IB1R. The effect of duration is
not manifested in the plots, in which surfaces have no slope except for bridge RB1R for small
intensities, where a steep surface with higher values of drift for longer duration is shown at low
seismic intensities. However, the surface does not show a constant slope while approaching
greater seismic intensities, indicating that for higher seismic intensities, duration does not play a
role. In addition, in this specific case the R2 value is small, which means that the surface
computed in (e) does not properly fit the results reflected by the points.

In terms of damage indices obtained, when referring to bridges IB1 and IB1R, it is important
to note that the results obtained followed two different approaches: (1) calculation of the global
damage index, where the contribution of each element is proportional to its energy dissipated by
yielding; and (2) the damage index of the column with the highest demands. Figure 4.13
indicates a slight tendency for the damage indicator DIP&A to increase when greater values of D5-
75 are observed, under the same level of spectral accelerations indicated by the slight inclination
of the surfaces. Figure 4.14 shows similar plots, but considering the Reinhorn & Valles damage
index as the response. In this case the effect of duration is more evident, but only for the cases in
which the bridge has not been retrofitted. In the irregular bridges IB1 and IBR1 both damage
index are naturally higher when considering the column that experienced the highest demands.

The results plotted above indicate that ground motion duration does not play a role for low
intensities of ground motion. However inspection of the surfaces at the highest ground motion
intensities, it can be noted that there is a tendency for greater values of damage indices.

4.6.3 Damage assessment based on incremental dynamic analysis


Figure 4.15 show the results in terms of IDA curves for peak deformation of bridges RB1
and RB1R, RB2 and RB2R and IB1 and IB1R, respectively. In these plots, the median peak
deformation observed is defined as the vector sum of peak deformations (square root of sum of
squares, SRSS) in both orthogonal directions of the bridges. The damage-index based IDA
curves are plotted in Figure 4.16: the first row shows results relative to Park & Ang index, while
the results shown in the second row are related with Reinhorn and Valles damage index.
121

Spectral Acceleration 4.0

3.0
Sa(T1), g

2.0

1.0

0.0
0.00 0.05 0.10 0.15 0.00 0.05 0.10 0.15 0.00 0.05 0.10 0.15
Peak Drift Ratio, Δ/H Peak Drift Ratio, Δ/H Peak Drift Ratio, Δ/H

(a) (b) (c)


Non-retrofitted - Short-duration Non-retrofitted - Long-duration
Retrofitted - Short-duration Retrofitted - Long-duration

Figure 4.15 - Median incremental dynamic analysis curves for vector sum of peak drift
ratios in X- and Y-directions based on results from bridge models: (a) RB1 and RB1R; (b) RB2
and RB2R; and (c) IB1 and IB1R.
122

0.8

0.6
DIP&A

0.4

0.2

0
0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0
Spectral Acceleration Spectral Acceleration Spectral Acceleration
Sa(T1), g Sa(T1), g Sa(T1), g

(a) (b) (c)

0.8

0.6
DIR&V

0.4

0.2

0.0
0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0
Spectral Acceleration Spectral Acceleration Spectral Acceleration
Sa(T1), g Sa(T1), g Sa(T1), g

(d) (e) (f)


Non-retrofitted - Short-duration Non-retrofitted - Long-duration
Retrofitted - Short-duration Retrofitted - Long-duration

Figure 4.16 – Median incremental dynamic analysis damage curves: (a) DIP&A curves for
RB1 and RB1R; (b) DIP&A curves for RB2 and RB2R; (c) DIP&A curves for IB1 and IB1R; (d)
DIR&V curves for RB1 and RB1R; (e) DIR&V curves for RB2 and RB2R; and (f) DIR&V curves for
IB1 and IB1R.

Figure 4.17 shows median IDA curves and fragility functions for bridges RB2, RB2R, IB1
and IB1R considering results obtained for short- and long-duration ground motions. The effects
of the retrofitting solution, ground motion duration and bridge irregularity can be analyzed in the
figure plots. Figure 4.17 (a) to (c) show the same curves plotted in Figure 4.15 and Figure 4.16,
but the curves relative to bridges RB2, RB2R, IB1 and IB1R are displayed in the same plot.
Figure 4.17 (d) to (f) display fragility functions that capture the probability of reaching a specific
value of drift or damage state (defined by reaching a specific value of damage index). In Figure
4.17 (d), the fragility functions represent the probability of reaching a threshold peak drift ratio
(pdr). The threshold values considered were different when considering non-retrofitted bridges
123

(RB2 and IB1) or retrofitted bridges (RB2R and IB1R). In the results shown for peak drift ratios,
a threshold value of 4% was considered for the non-retrofitted bridges, while 10% was assumed
for the retrofitted cases. The peak drift ratios considered are not indicative of bridge collapse,
however, when comparing pdr-based fragility functions with damage-index fragility functions,
an evident similarity is observed. More experimental tests are required to be able to define values
for collapse based on the columns peak drift ratio, considering the type of retrofit solution
applied. As an example, the results from cyclic tests performed by Lostra (2016) led to drift
ratios greater than 8% without any significant damage observed.

Figure 4.17 (e) and (f) plot the damage index-based fragility functions that compute the
probability of reaching the collapse limit state. The “collapse” qualification is based on a damage
index greater than 0.8 as suggested in Ang et al. (Ang et al. 1993). Despite only curves with the
probability of collapse of bridges RB2, RB2R, IB1, and IB1R being shown, the probability of
reaching the moderate and severe damage states (according to Ang et al. 1993) together with the
probability of reaching collapse were determined for the intensities Sa(T1) = 0.5g, 1.0g and 1.5g
for all bridges studied, and are shown in Table 4.1, Table 4.2 and Table 4.3.
124

4.0 0.8 0.8


Spectral Acceleration

3.0 0.6 0.6

DIR&V
Sa(T1), g

DIP&A
2.0 0.4 0.4

1.0 0.2 0.2

0.0 0.0 0.0


0 0.05 0.1 0.15 0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0
Peak Drift Ratio Δ/H Spectral Acceleration Spectral Acceleration
Sa(T1), g Sa(T1), g

(a) (b) (c)

1.0 1.0 1.0


pdr = 4%
P [Peak Drift Ratio> pdr]

0.8 0.8 0.8


P [DIP&A > 0.8]

P [DIR&V > 0.8]


0.6 0.6 0.6

0.4 0.4 0.4

0.2 pdr = 10% 0.2 0.2

0.0 0.0 0.0


0.0 1.0 2.0 3.0 4.0 0.0 1.0 2.0 3.0 4.0 0.0 1.0 2.0 3.0 4.0
Spectral Acceleration Spectral Acceleration Spectral Acceleration
Sa(T1), g Sa(T1) g Sa(T1) g

(d) (e) (f)

Bridge RB2R Bridge RB2R Bridge RB2 Bridge RB2


Series1 Series2 Series3 Series4
Short-duration Long-duration Short-duration Long-duration
Bridge IB1R
Series5 Bridge IB1R
Series6 Bridge IB1
Series13 Bridge IB1
Series14
Short-duration Long-duration Short-duration Long-duration

Figure 4.17 - (a) Median IDA curves computing the peak vector sum of drift ratios in X- and
Y-directions; (b) Evolution of median DIP&A curves with the increase of seismic intensities; (c)
Evolution of median DIR&V curves with the increase of seismic intensities; (d) Fragility curves
that compute probability of exceeding a determined drift ratio (4% for the non-retrofitted bridges
and 10% for the retrofitted bridges); (e) Fragility curves that compute probability of reaching
collapse (DI > 0.8) based on the Park and Ang damage index DIP&A; and (f) Fragility curves that
compute probability of reaching collapse (DI>0.8) based on the Reinhorn and Valles damage
index DIR&V
125

4.7 Discussion

4.7.1 Effect of bridge retrofits using TiABs


When looking specifically to the effect of the retrofit solution adopted for the columns of
these bridges, an evident increase is observed in terms of deformation capacities for the
retrofitted bridges. From the IDA curves observed in Figure 4.15 it can be seen that the
performance of the retrofitted bridges RB1R, RB2R and IB1R is enhanced, with an increase of
capacity of approximately twice the non-retrofitted bridge models. The same plots also show that
bridge RB1 has capacity to deform up to 5% drift ratio and handle shaking intensities of Sa(T1) =
0.75g. After applying the retrofit, bridge RB1R shows capacity to deform up to 10% drift ratio at
Sa(T1) = 1.5g. Bridges RB2 and IB1 with lower fundamental periods have more strength and
deformation capacities reaching values around 7% of peak drift ratio and Sa(T1) of 1.5g. The
retrofitted bridge RB2Rreaches intensities greater than Sa(T1) of 3.0g for values of peak drift
ratio of 10%. Similar increases in capacity can be observed for the IB1R bridge model results,
regarding peak drift ratios.

From observation of the results plotted in Figure 4.16, the damage indices for the non-
retrofitted bridges, indicate structural collapse at: (1) Sa(T1) of approximately 0.5g for RB1; (2)
Sa(T1) ≈ 0.8g for RB2, and (3) Sa(T1) ≈ 0.75g for IB1. For the retrofitted bridge models the
collapses are observed at: (1) Sa(T1) ≈ 1.5g for RB1R, (b) Sa(T1) > 2.0g for RB2R, and (3)
Sa(T1) > 1.2g for IB1R. In summary, the results highlight the great effect of the retrofit solution
on the structural capacity increasing their capacity 2 to 3 times independently of the damage
indicator used. Table 4.1, Table 4.2 and Table 4.3 show directly the consequences of retrofitting
the bridge columns by computing the probability of reaching a specific damage level, showing
that the probability of reaching any damage states reduces significantly when retrofitted,
considering medium and high intensities.
126

Table 4.1 - Reduction of probability of reaching a specific damage state based on DIP&A and
DIR&V when retrofitted, for a ground motion intensity of Sa(T1) = 0.5g

Bridge B1 → Bridge B1R Bridge B2 → Bridge B2R


Sa(T1) = 0.5g
Short-duration Long-duration Short-duration Long-duration
DIP&A
Moderate Damage 91% → 50% 100% → 40% 68% → 9% 86% → 3%
Severe Damage 80% → 35% 96% → 16% 47% → 5% 48% → 1%
Collapse 53% → 20% 46% → 5% 21% → 2% 11% → 0%
DIR&V
Moderate Damage 80% → 45% 95% → 30% 51% → 7% 53% → 2%
Severe Damage 70% → 32% 80% → 15% 36% → 4% 32% → 1%
Collapse 52% → 20% 49% → 7% 22% → 2% 17% → 0%
Bridge IB1 {IB2}* [IB3]* → Bridge IB1R {IB2R}*[IB3R]*
Short-duration Long-duration
DIP&A
Moderate Damage 73% → 21% 95% {96%} [96%]→ 7%
Severe Damage 52% {53%} [53%]→11% 59% {60%} [61%]→ 2%
Collapse 23% {24%} [24%]→ 4% 14% [15%]→ 1%
DIR&V
Moderate Damage 58% {59%} [59%] → 19% {18%} 71% {72%} [73%]→ 6%
Severe Damage 44% {45%} [45%]→ 11% 48% {49%} [50%]→ 2%
Collapse 27% → 5% 27% [29%]→ 2%
* only shown if different value is obtained
127

Table 4.2 - Reduction of probability of reaching a specific damage state based on DIP&A and
DIR&V when retrofitted, for a ground motion intensity of Sa(T1) = 1.0g

Bridge B1 → Bridge B1R Bridge B2 → Bridge B2R


Sa(T1) =1.0g
Short-duration Long-duration Short-duration Long-duration
DIP&A
Moderate Damage 99% → 79% 100% → 93% 94% → 38% 100% → 35%
Severe Damage 95% → 64% 95% → 70% 84% → 23% 98% → 17%
Collapse 80% → 42% 89% → 31% 55% → 11% 60% → 5%
DIR&V
Moderate Damage 94% → 73% 100% → 87% 87% → 32% 98% → 28%
Severe Damage 91% → 61% 99% → 67% 75% → 21% 89% → 16%
Collapse 79% → 44% 90% → 41% 55% → 12% 67% → 8%
Bridge IB1 {IB2}* [IB3]* → Bridge IB1R {IB2R}* [IB3R]*
Short-duration Long-duration
DIP&A
Moderate Damage 96% → 61% 100% → 73% [75%]
Severe Damage 88% →41% 100% → 40% [41%]
Collapse 62% → 19% 73% {72%} [74%] → 12%
DIR&V
Moderate Damage 90% → 57% 100% → 66% [67%]
Severe Damage 83% → 40% 97% [98%] → 43% [44%]
Collapse 65% → 23% 85% {86%} [86%]→ 25%
* only shown if different value is obtained
128

Table 4.3 - Reduction of probability of reaching a specific damage state based on DIP&A and
DIR&V when retrofitted, for a ground motion intensity of Sa(T1) = 1.5g

Bridge B1 → Bridge B1R Bridge B2 → Bridge B2R


Sa(T1) = 1.5g
Short-duration Long-duration Short-duration Long-duration
DIP&A
Moderate Damage 100% → 90% 100% → 99% 99% → 61% 100% → 68%
Severe Damage 98% → 79% 100% → 92% 95% → 42% 100% → 43%
Collapse 90% → 57% 90% → 57% 75% → 22% 87% → 16%
DIR&V
Moderate Damage 98% → 85% 100% → 98% 96% → 55% 100% → 60%
Severe Damage 96% → 77% 100% → 90% 90% → 40% 99% → 41%
Collapse 89% → 60% 98% → 71% 75% → 25% 89% → 25%
Bridge IB1 {IB2}* [IB3]* → Bridge IB1R {IB2R}* [IB3R]*
Short-duration Long-duration
DIP&A
Moderate Damage 99% →82% 100% → 97%
Severe Damage 96% [97%]→63% [64%] 100% → 79% [80%]
Collapse 82% [81%]→ 37% 95% {94%} → 36%
DIR&V
Moderate Damage 97% → 78% [79%] 100% → 94% [95%]
Severe Damage 94% → 62% [63%] 100% → 81% [82%]
Collapse 84% → 43% 98% {97%} → 58%
* only shown if different value is obtained

4.7.2 Effect of bridge irregularity


The plots overlaying results for bridges RB2, RB2R, IB1 and IB1R in Figure 4.17 allow for
the assessment of the effect of bridge irregularity on the four-span bridges. In terms of median
peak drift ratio and damage index curves obtained from IDA, negligible differences are observed
for the non-retrofitted bridges RB2 and IB1. On the other hand, in the results obtained for the
retrofitted bridges RB2R and IB1R, the bridge irregularity plays an important role in the
enhancement provided by the retrofit solution. The capacity of bridge RB2R is significantly
greater when compared with IB1R as can be observed from all plots of Figure 4.17, which lead
to the fact that in bridges with irregular geometry, the retrofit solution is not able to provide a
sufficient enhancement of the performance of the short columns when subjected to higher ground
motion shaking intensities, which is expected due to the increased demand in the columns.
129

4.7.3 Ground motion duration effects


The effects of ground shaking duration are not observed when analyzing IDA curves for the
peak drift ratio obtained for any bridge studied, being consistent with the results shown in .

The median IDA curves based on DIP&A do not capture effect of ground motion duration on
bridges RB1, RB1R, and RB2, while a slight effect can be observed in the values of damage
indices obtained for bridges RB2R, IB1 and IB1R when subjected to greater ground motion
intensities, but only for level of damage index greater than 0.4.

When considering the DIR&V, no discernable difference is observed between short- and long-
duration results for bridge RB1, slight effects are captured for bridges RB1R, RB2, and IB1 at
levels of damage exceeding the severe damage limit state, and significant effects of duration are
detected in the retrofitted bridge models RB2R and IB1R for levels of damage exceeding the
moderate damage limit state. At this level of damage, a difference in peak capacity of 0.5g is
observed, which translated to an impact of ground motion duration of 30% at the collapse limit
state for bridge models RB2R and IB1R.

From observation of fragility functions in Figure 4.17 (d), the effect of ground motion is
only captured for higher ground motion intensities. This trend is followed by the curves that
compute the probability of reaching collapse, and confirmed when comparing probabilities for
Sa(T1) of 0.5g, 1.0g and 1.5g relative to short- and long-duration shown in Table 4.1, Table 4.2
and Table 4.3. On the other hand, if the focus is on other damage states, higher probabilities of
reaching moderate and severe damage states are observed when subjected by long-duration
ground motions for smaller intensities as can be seen in Table 4.1 with Sa(T1) = 0.5g.

4.8 Conclusions

• The effect of ground motion duration was evaluated on the response of models of typical
Oregon bridges built between 1950 and 1970, with and without seismic retrofitting. The
bridges analyzed are based on the already demolished Mackenzie River on Highway
Interstate 5, a 3-span regular bridge with median dimensions obtained from the Oregon
Department of Transportation’s database of bridges, and three irregular bridges with three
different column heights, including one column with the 16th percentile of column height,
130

one with the median height, and one with the 84th percentile of column height considering
all bridges in Oregon built during the mentioned period. A novel retrofitting solution
reported recently in Lostra (2016) was considered to provide enhanced cyclic
performance of the bridge columns, which in all five bridges had lap-splices with
insufficient length to be able to adequately embed and transfer forces from the column
bars to the footing starter bars. Thus, the bridges considered can be classified in two
groups, including those with: (1) as-designed, and (2) columns retrofitted with vertical
TiABs embedded into epoxy-filled drilled holes in the footings and a concrete shell
encased by a TiAB continuous spirals. Thus, in total, ten bridges were studied.
Models of all non-retrofitted and retrofitted bridges were developed in OpenSees with
special attention provided to the modeling of the lap-splice behavior, and to the modeling
of titanium bars and the effects of confinement provided by the titanium spiral encased
concrete shell. Cyclic analyses were performed to the column models and validated with
results obtained from laboratory tests of real scale specimens of columns representative
of typical Oregon bridges.
Results of the damage assessment performed for all bridges are based on peak drift ratios
and on the damage indices by Park & Ang (DIP&A) and Reinhorn & Valles (DIR&V) and
led to the following conclusions:
• First, the statistical analysis of the results of the bridge models when subjected to the set
of unscaled short-duration ground motions and the unscaled long duration ground
motions indicate that:
o Negligible effects of ground motion duration on these bridges were observed for
both non-retrofitted and retrofitted bridges, showing slightly greater values of the
damage indices at greater significant durations and at higher values of Sa(T1),
more evident when evaluated the DIR&V for all intensities of ground shaking.
o The effects of increased ground motion duration are more significant for irregular
bridges.
• The TiABs used as retrofitting solution provide an increase of 2 to 3 times to the bridge
structural capacities as measured from the IDA curves.
131

• The fragility functions for the retrofitted bridges and for exceeding a limit state of a peak
drift ratio of 10% are identical to the ones developed based on the damage indices for the
collapse limit state corresponding to a damage index threshold of 0.8.
• The bridge irregularity does not play an important role when estimating the strength and
deformation capacity of the non-retrofitted bridges. However, the retrofit solution is
found 1.5 to 2 times more efficient in terms of capacity provided, when applied to regular
bridges compared with irregular bridges, demonstrating the important role played by the
bridge irregularity on the impact of the retrofit solution applied to the bridge columns.

The following points should be addressed in future works:

• Model different abutments topologies, including those with buried piers. The columns
would work as piles, therefore a soil-pile interaction modeling should be also considered
for such abutments.
• Perform cyclic testes on TiABs to confirm and enhance the titanium cyclic model that
was proposed in this work
132

4.9 References

Abedi, K., Afshin, H., and Shirazi, M. R. N. (2010). “Numerical study on the seismic retrofitting
of reinforced concrete columns using rectified steel jackets.” Asian Journal of Civil
Engineering (Building and Housing), 11(2), 219–240.

ACI. (2011). Building Code Requirements for Structural Concrete (ACI 318-11). American
Concrete Institute, Farmington Hills, MI.

Ang, A. H.-S., Kim, W. J., and Kim, S. B. (1993). “Damage Estimation of Existing Bridge
Structures.” Structural Engineering in Natural Hazards Mitigation:ASCE Structures
Congress, Irvine, CA, 1137–1142.

Barbosa, A. R., Ribeiro, F. L. A., and Neves, L. C. (2017). “Influence of earthquake ground-
motion duration on damage estimation: application to steel moment resisting frames.”
Earthquake Engineering and Structural Dynamics, 46(1), 27–49.

Bazaez, R., and Dusicka, P. (2016a). “Cyclic Loading for RC Bridge Columns Considering
Subduction Megathrust Earthquakes.” Journal of Bridge Engineering, American Society of
Civil Engineers, 21(5), 4016009.

Bazaez, R., and Dusicka, P. (2016b). “Cyclic behavior of reinforced concrete bridge bent
retrofitted with buckling restrained braces.” Engineering Structures, Elsevier, 119, 34–48.

Belejo, A., Barbosa, A. R., and Bento, R. (2017). “Influence of Ground Motion Duration on
Damage Index-based Fragility Assessment of a Plan-Asymmetric Non-Ductile Reinforced
Concrete Builing.” Engineering Structures, 151, 682–703.

Bojorquez, E., Iervolino, I., Manfredi, G., and Cosenza, E. (2006). “Influence of ground motion
duration on degrading SDOF systems.” Proceedings of 1st European Conference on
Earthquake Engineering and Seismology, Geneve, Switzerland.

Bommer, J. J., Magenes, G., Hancock, J., and Penazzo, P. (2004). “The Influence of Strong-
Motion Duration on the Seismic Response of Masonry Structures.” Bulletin of Earthquake
Engineering, 2(1), 1–26.

Cairns, J., and Arthur, P. D. (1979). “Strength of lapped splices in reinforced concrete columns.”
Journal Proceedings, 76(2), 277–296.

Chai, Y. H. (2005). “Incorporating low-cycle fatigue model into duration-dependent inelastic


design spectra.” Earthquake Engineering and Structural Dynamics, 34(1), 83–96.
133

Chandramohan, R., Baker, J. W., and Deierlein, G. G. (2016). “Quantifying the Influence of
Ground Motion Duration on Structural Collapse Capacity Using Spectrally Equivalent
Records.” Earthquake Spectra, 32(2), 927–950.

Chang, S.-Y., Li, Y.-F., and Loh, C.-H. (2004). “Experimental study of seismic behaviors of as-
built and carbon fiber reinforced plastics repaired reinforced concrete bridge columns.”
Journal of Bridge Engineering, American Society of Civil Engineers, 9(4), 391–402.

Chowdhury, S. R., and Orakcal, K. (2012). “An analytical model for reinforced concrete
columns with lap splices.” Engineering Structures, Elsevier, 43, 180–193.

Cornell, A. C. (1997). “Does duration really matter?” Proceedings of the FHWA/NCEER


Workshop on the National Representation of Seismic Ground Motion for New and Existing
Highway Facilities, Burlingame, Calif., organized by NCEER (project 106-F-5.4.1) and
ATC (project ATC-18-1), 125–133.

Douglas, D. J., and Davis, E. H. (1964). “The movement of buried footings due to moment and
horizontal load and the movement of anchor plates.” Geotechnique, Thomas Telford Ltd,
14(2), 115–132.

Duncan, J. M., and Mokwa, R. L. (2001). “Passive earth pressures: theories and tests.” Journal of
Geotechnical and Geoenvironmental Engineering, American Society of Civil Engineers,
127(3), 248–257.

Dusicka, P., and Lopez, A. (2016). Impact of Cascadia Subduction Zone Earthquake on the
Seismic Evaluation Criteria of Bridges. Portland.

Dutta, A., and Mander, J. B. (2001). “Energy Based Methodology for Ductile Design pf
Concrete Columns.” ASCE Journal of Structural Engineering, 127(12), 1374–1381.

Earth Mechanics, I. N. C. (2005). Field investigation report for abutment backfill


characterization. University of California, San Diego, CA.

ElGawady, M., Endeshaw, M., McLean, D., and Sack, R. (2010). “Retrofitting of Rectangular
Columns with Deficient Lap Splices.” Journal of Composites for Construction, 14(1), 22–
35.

Endeshaw, M. a., ElGawady, M., Sack, R. L., and McLean, D. I. (2008). Retrofit of Rectangular
Bridge Columns using CFRP Wrapping.

Feng, M. Q., Bhatia, H., Baird, K., and Elsanadedy, H. (1999). Structural Qualification Testing
of Composite-jacketed Circular and Rectangular Bridge Columns. Irvine, CA.
134

Filippou, F. C., Popov, E. P., and Bertero, V. V. (1983). Effects of Bond Deterioration on
Hysteretic Behaviour of Reinforced Concrete Joints. Report to the National Science
Foundation. Earthquake Engineering Research Center, Berkeley, California.

Gidaris, I., Padgett, J. E., Barbosa, A. R., Chen, S., Cox, D., Webb, B., and Cerato, A. (2016).
“Multiple-Hazard Fragility and Restoration Models of Highway Bridges for Regional Risk
and Resilience Assessment in the United States: State-of-the-Art Review.” Journal of
Structural Engineering, 04016188.

Girard, C., and Bastien, J. (2002). “Finite-Element Bond-Slip Model for Concrete Columns
under Cyclic Loads.” Journal of Structural Engineering, 128(12), 1502–1510.

Hancock, J., and Bommer, J. J. (2006). “A state-of-knowledge review of the influence of strong-
motion duration on structural damage.” Earthquake Spectra, 22(3), 827–845.

Harajli, M. H. (2009). “Bond stress--slip model for steel bars in unconfined or steel, FRC, or
FRP confined concrete under cyclic loading.” Journal of Structural Engineering, American
Society of Civil Engineers, 135(5), 509–518.

Harajli, M. H., Hamad, B. S., and Rteil, A. A. (2004). “Effect of Confinement of Bond Strength
between Steel.” ACI Structural Journal, 101(5).

Haselton, C. B., Liel, A. B., Lange, S. T., and Deirlein, G. (2008). Beam-Column Element Model
Calibrated for Predicting Flexural Response Leading to Global Collapse of RC Frame
Buildings. Berkeley, California.

Hoshikuma, J., Kawashima, K., Nagaya, K., and Taylor, A. W. (1997). “Stress-strain model for
confined reinforced concrete in bridge piers.” Journal of Structural Engineering, American
Society of Civil Engineers, 123(5), 624–633.

Iacobucci, R. D., Sheikh, S. A., and Bayrak, O. (2003). “Retrofit of square concrete columns
with carbon fiber-reinforced polymer for seismic resistance.” ACI Structural Journal,
American Concrete Institute, 100(6), 785–794.

Kappos, A. J., Manolis, G. D., and Moschonas, I. F. (2002). “Seismic assessment and design of
R/C bridges with irregular congiguration, including SSI effects.” Engineering Structures,
24(10), 1337–1348.

Katsumata, H., Kimura, K., and Kobatake, Y. (1998). “Seismic Retrofitting Technique Using
Carbon Fibers For Reinforced Concrete Buildings.” Strain, 2(4), 0.

Kayen, R. E., and Mitchell, J. K. (1997). “Assessment of liquefaction potential during


earthquakes by Arias intensity.” Journal of Geotechnical and Geoenvironmental
Engineering, American Society of Civil Engineers, 123(12), 1162–1174.
135

Kim, S.-H., and Shinozuka, M. (2004). “Development of fragility curves of bridges retrofitted by
column jacketing.” Probabilistic Engineering Mechanics, Elsevier, 19(1), 105–112.

Kramer, S. L., and Mitchell, R. A. (2006). “Ground motion intensity measures for liquefaction
hazard evaluation.” Earthquake Spectra, 22(2), 413–438.

Lostra, M. (2016). “Seismic Performance of Square Reinforced Concrete Columns Retrofitted


with Titanium Alloy Bars.” Oregon State University.

Lostra, M., Higgins, C., and Barbosa, A. R. (2017). “Seismic retrofit of reinforced concrete
bridge columns using titanium-alloy bars.” Portuguese Journal of Structural Engineering,
3, 75–82.

Lowes, L. N., Mitra, N., and Altoontash, A. (2003). A beam-column joint model for simulating
the earthquake response of reinforced concrete frames. Pacific Earthquake Engineering
Research Center, College of Engineering, University of California Berkeley.

Lukose, K., Gergely, P., and White, R. N. (1982). “Behavior of reinforced concrete lapped
splices for inelastic cyclic loading.” Journal Proceedings, 355–365.

Mackie, K. R., and Stojadinović, B. (2005). Fragility basis for California highway overpass
bridge seismic decision making. Pacific Earthquake Engineering Research Center, College
of Engineering, University of California, Berkeley, Berkeley, California.

Mackie, K. R., and Stojadinović, B. (2006). “Post-earthquake functionality of highway overpass


bridges.” Earthquake engineering and structural dynamics, Wiley Online Library, 35(1),
77–93.

Malley, J. O., Deierlein, G., Krawinkler, H., Maffei, J. R., Pourzanjani, M., Wallace, J., and
Heintz, J. (2010). Modeling and acceptance criteria for seismic design and analysis of tall
buildings. Technical Report PEER/ATC 72-1. Redwood City.

Mander, J. B., Priestley, M. J. N., and Park, R. (1988). “Theoretical stress-strain model for
confined concrete.” Journal of structural engineering, American Society of Civil Engineers,
114(8), 1804–1826.

Manfredi, G., Polese, M., and Cosenza, E. (2003). “Cumulative demand of the earthquake
ground motions in the near source.” Earthquake Engineering and Structural Dynamics,
32(12), 1853–1865.

McLean, D., and Bernards, L. (1992). Seismic Retrofitting of Rectangular Bridges Column for
Shear. Pullman, Washington.

Melek, M., and Wallace, J. W. (2004). “Cyclic behavior of columns with short lap splices.” ACI
Structural Journal, 101(6), 802–811.
136

Menegoto, M., and Pinto, P. E. (1973). “Menegoto-Method of analysis for cyclically loaded R.C.
plane frames including changes in Geometry and non-elastic behavior of elements under
combined normal force and bending.” Proceedings of the symposium resitance and
Ultimate Deformability of Structures Acted on by Well Repeated Loads, Lisboa.

Mohammadi-Tamanani, M., Gian, Y., and Ayoub, A. (2014). “Design of bridges with unequal
pier heights.” Structures Congress 2014, Boston, Massachusetts, 677–686.

Nassar, A., and Krawinkler, H. (1991). Seismic demands for SDOF and MDOF systems. No. 95.
John A. Blume Earthquake Engineering Center, Department of Civil Engineering. John A.
Blume Earthquake Engineering Center, Department of Civil Engineering, Standford,
California.

Nielson, B. G., and DesRoches, R. (2007). “Seismic fragility methodology for highway bridges
using a component level approach.” Earthquake Engineering & Structural Dynamics, 36,
823–839675.

Oyarzo-vera, C., and Chouw, N. (2008a). “Effect of earthquake duration and sequences of
ground motions on structural responses.” Proceedings of the 10th International Symposium
on Structural Engineering of Young Experts, Changsha, Scottland.

Oyarzo-vera, C., and Chouw, N. (2008b). “Effect of earthquake duration and sequences of
ground motions on structural responses.” Proceedings of the 10th International Symposium
on Structural Engineering of Young Experts.

Padgett, J. E., and DesRoches, R. (2009). “Retrofitted bridge fragility analysis for typical classes
of multispan bridges.” Earthquake Spectra, 25(1), 117–141.

Park, Y. J., Ang, A. H. S., and Wen, Y. K. (1987). “Damage-Limiting Aseismic Design of
Buildings.” Earthquake Spectra.

Paulay, T. (1982). “Lapped splices in earthquake-resisting columns.” Journal Proceedings, 458–


469.

Raghunandan, M., and Liel, A. B. (2013). “Effect of ground motion duration on earthquake-
induced structural collapse.” Structural Safety, Elsevier, 41, 119–133.

Raghunandan, M., Liel, A. B., and Luco, N. (2015). “Collapse Risk of Buildings in the Pacific
Northwest Region due to Subduction Earthquakes.” Earthquake Spectra, 31(4), 2087–2115.

Ruiz-García, J. (2010). “On the influence of strong-ground motion duration on residual


displacement demands.” Earthquake and Structures, 1(4), 327–344.

Schlick, B. M., and Breña, S. F. (2004). “Seismic rehabilitation of reinforced concrete bridge
columns in moderate earthquake regions using FRP composites.” Proc. of the 13th World
137

Conference on Earthquake Engineering,(CD-Rom, paper n. 508), Vancouver, BC, Canada,


August, 1–6.

Scott, B. D., Park, R., and Priestley, M. J. N. (1982). “Stress-Strain Behavior of Concrete
Confined by Overlapping Hoops at Low and High Strain Ratio Rates.” Journal of American
Concrete Institute, 79(6), 496–498.

Shome, N., Cornell, C. A., and Bazurro, P. (1998). “Earthquakes, records, and nonlinear
responses.” Earthquake Spectra, 14(3), 469–500.

Sideras, S. S., and Kramer, S. L. (2012). “Potential implications of long duration ground motions
on the response of liquefiable soil deposits.” 9th International Conference on Urban
Earthquake Engineering/4th Asia Conference on Earthquake Engineering, Tokyo, Japan.

Song, R., Li, Y., and van de Lindt, J. W. (2014). “Impact of earthquake ground motion
characteristics on collapse risk of post-mainshock buildings considering aftershocks.”
Engineering Structures, Elsevier Ltd, 81, 349–361.

Stevens, N. J., Uzumeri, S. M., Will, G. T., and others. (1991). “Constitutive model for
reinforced concrete finite element analysis.” Structural Journal, 88(1), 49–59.

Tremblay, R., and Atkinson, G. M. (2001). “Comparative Study of the Inelastic Seismic Demand
of Eastern and Western Canadian Sites.” Eartquake Spectra, 17(2), 333–358.

Wilson, P., and Elgamal, A. (2006). “Large scale measurement of lateral earth pressure on bridge
abutment back-wall subjected to static and dynamic loading.” Proceedings of New Zealand
Workshop on Geotechnical Earthquake Engineering, Christchurch, New Zealand, 307–315.

Yassin, M. H. M. (1994). Nonlinear analysis of prestressed concrete sructures under monotonic


and cyclic loads. University of California, Berkeley.

Zhao, J., and Sritharan, S. (2007). “Modeling of strain penetration effects in fiber-based analysis
of reinforced concrete structures.” ACI structural journal, American Concrete Institute,
104(2), 133.
138

5 INFLUENCE OF GROUND-MOTION DURATION ON DAMAGE ESTIMATION


OF STEEL MOMENT RESISTING FRAMES ACCOUNTING FOR SOIL-
STRUCTURE INTERACTION EFFECTS

Andre Belejo, Andre Ramos Barbosa, Jonathan Huffman, and Armin Stuedlein

EQUATION CHAPTER 5 SECTION 1

Journal of Structural Engineering


American Society of Civil Engineers
1801 Alexander Bell Drive, Reston, VA 20191-4400
Prepared for Submission
139

5.1 Overview

When addressing the topic of ground motion duration, it is accepted by the engineering
community that earthquake strong ground motion duration affects the response of soils. In
addition, the soil-structure interaction (SSI) effects on the built environment have also been
studied. Recent publications have highlighted the importance of ground motion duration on the
estimation of the seismic response and seismic damage of building and bridges structures, where
it has been reported that for strong shaking intensities, buildings and bridges tend to exhibit
increased damage with the increasing duration of the ground motions. However, limited research
has addressed the combined effects of earthquake strong ground motion duration and SSI. Thus,
the objective of this study is to improve the understanding of the combined effects of earthquake
strong ground motion duration and SSI on probabilistic structural damage. Two steel moment
resisting frame buildings supported on shallow foundations are used as examples. A new semi-
empirically phenomenological Q-y type multi-spring model is implemented for use in the
modeling of shallow foundations. Modeling uncertainty and spatial variability of the soil
parameters are considered and incorporated into the nonlinear finite element models. Spatial
variability is modeled using random field theory. Seismic fragility curves are developed for the
cases with fixed-base models and for the cases considering SSI effects. The response and
damage parameters considered in developing the fragility functions include peak drift ratios,
residual drift ratios, and two commonly used damage indices. Capacity estimation of the
buildings is based on incremental dynamic analyses (IDA). The inclusion of SSI reduces the base
shear in the building models, while increases the peak and residual interstory drift ratios. In
addition, results in this chapter confirm that ground motion duration plays an important role in
the damage prediction by considering SSI effects. Therefore, it is suggested that structural
capacities may be reduced to account for the effects of increased ground motion duration, in
buildings design/assessment guidelines.
140

5.2 Introduction

When considering the ground motion duration on the structural response of civil infra-
structure, several studies identified that ground motion duration plays an important role: (1)
Chandramohan et al. (2016) and Barbosa et al. (2017) showed that for the same intensity of
shaking, characterized by having the same response spectral accelerations, longer duration
motions induce higher peak interstory drift ratios; (2) increased ground motion duration induced
more damage accumulation (Dutta and Mander 2001; Manfredi et al. 2003; Oyarzo-vera and
Chouw 2008; Barbosa et al. 2017; Belejo et al. 2017); (3) Ruiz-García (2010) indicated that
longer duration ground motions are responsible for greater residual deformations, and (4) long
duration ground motions cause higher probabilities of structural collapse when compared with
short-duration ground motions (Raghunandan and Liel 2013; Song et al. 2014; Raghunandan et
al. 2015; Chandramohan et al. 2016). However, these studies neglected the effects of soil-
structure interaction.

Among the analytical studies mentioned above, the effect of soil-structure interaction (SSI)
were not considered, even though it is well known that SSI effects can be both beneficial or
detrimental on the magnitude of peak force and deformation demands on structures (Mylonakis
and Gazetas 2000; Gazetas and Mylonakis 2001; Stewart et al. 2003; Gazetas and Apostolou
2004; Pecker and Chatzigogos 2010; Moghaddasi et al. 2010, 2011, 2012; Toh et al. 2011).

Considering the buildings supported on shallow foundations, the most common approaches
used on modeling SSI effects are: continuum finite element and boundary element approaches,
macro-element formulations, and Winkler-based approaches. The continuum approach assumes a
semi-infinite and isotropic material able to consider multiple soil layers, however involves
extensive computational effort (Drucker and Prager, 1952; Fletcher and Hermann, 1971; Dutta
and Roy, 2002). Macro-element models consist of single element models, in which the soil-
structure interaction response is captured using plasticity-based formulations. Cremer et al.
(2001), Gajan (2006), Paolucci et al. (2008) and Figini et al. (2012) developed macro-element
models for shallow foundations. The main advantage with respect to other approaches is
characterized by the coupling between all the macro-element DOFs. The Winkler-based
modeling approach has been one of the most applied by the research community due to its
141

efficiency and simplicity, and thus is the model used also in this study. Bartlett (1976) and
Taylor et al. (1980) predicted moment-rotation behavior of rigid footings using elastic-perfectly-
plastic compression only springs, coupled with Coulomb friction slider elements that captured
the uplifting of the foundation. Chopra and Yim (1985) developed simplified expressions to
determine the base shear resistance of flexible structures allowed to uplift, showing impact of the
uplift and foundation flexibility on the fundamental mode of vibration, but little effect on higher
modes. Nakaki and Hart (1987) used Winkler springs with zero tension capacity and elastic
compressive resistance, together with viscous dampers at the base of a shear wall structure.
Results indicated that the fundamental period of the structure increases substantially resulting in
greater ductility demands on the structure due to the induced foundation uplift. Allotey and El
Naggar (2008) adopted their previous Winkler-based modeling concept (Allotey and El Naggar
2003) for capturing the cyclic response of shallow foundations, considering linear backbone
curves to predict the moment–rotation and settlement response. Raychowdhury (2008) developed
a numerical model based on the Beam-on-Nonlinear-Winkler-Foundation (BNWF) to capture the
soil-foundation behavior using shallow foundations, modifying backbone curves typically used
for modeling soil-pile response (Boulanger et al. 1999). The model can reasonably predict
experimentally measured footing response in terms of moment, shear, settlement, and rotational
demands as well as the general hysteresis shape of the moment–rotation, settlement–rotation and
shear–sliding curves. Gajan et al. (2008; 2010) developed a contact interface model (CIM),
providing constitutive relations between cyclic loads and displacements at the footing-soil
interface of a shallow rigid foundation subjected to combined moment, shear, and axial loading.
Harden and Hutchingson (2009) adopted a BNWF simulation methodology for modeling shallow
foundation-structure systems, where seismically-induced rocking plays a predominant role in
their response, finding a reasonable correlation between results of the nonlinear Winkler-based
approach and experimental tests, when appropriate and calibrated material model and soil
properties were defined. Raychowdhury (2011) applied nonlinear SSI models on low-rise SMRF
buildings and observed that forces are reduced and displacement demands increase even though
interstory drift ratios were reduced, when accounting for the foundation nonlinearity as
compared with the results obtained from fixed-base and elastic SSI models.
142

Before performing seismic analyses of soil-structure systems, it is important to ensure that


the foundations have been designed for both serviceability and ultimate limit states, before
subjecting the system to ground motion shaking. Focusing on serviceability limit state, Huffman
et al. (2015) proposed a simple reliability-based design (RBD) procedure for assessing the
allowable immediate displacement of a spread footing supported on plastic soils, considering a
desired serviceability limit state. A bivariate normalized bearing pressure–displacement model to
estimate the mobilized resistance associated with a given displacement was developed and
calibrated using a high quality database of full-scale loading tests. The loading test data,
compiled by Strahler (2012) and Strahler and Stuedlein (2014), was used to characterize the
uncertainty associated with the model and incorporated into an appropriate reliability-based
performance function. This uncertainty is characterized by differing geologic conditions,
construction techniques, and other variables in Huffman et al. (2015). Monte Carlo simulations
were then used to calibrate a resistance factor with consideration of the uncertainty in the bearing
pressure–displacement model, bearing capacity, applied bearing pressure, allowable
displacement, and footing width. This model is valid also up to large displacements, which can
induce exceedance of ultimate limit states.

The work presented herein extends the work developed in Barbosa et al. (2017). Barbosa et
al. (2017) evaluated the influence of ground motion duration on the structural damage of 3-story
and 9-story steel moment resisting frame (SMRF) buildings using two-dimensional nonlinear
finite element fixed-base models. A damage assessment was performed using damage indices
proposed by Park and Ang (1985) DIP&A, and Reinhorn and Valles (1995) DIR&V, which are
damage metrics that combine peak deformations with the effects of cumulative loading and
inelastic measures of the response to account for low-cycle fatigue. Results in Barbosa et al.
(2017) indicated that increased duration of the ground motions induced larger peak interstory
drift ratios (IDRs) and larger damage indices for greater spectral accelerations. Results indicated
that performance-based and code-based assessment methodologies should be revised to consider
damage measures that are sensitive to ground motion duration, since the current assessment
methodologies do not capture these effects.
143

The main objective of this chapter is to develop an understanding of how the combined
effects of ground motion duration and SSI impact the structural performance of the two pre-
Northridge steel MRF buildings analyzed in Barbosa et al. (2017). This is achieved by extending
the Huffman et al. (2015) approach for modeling foundations of the buildings to capture cyclic
response, and by developing and implementing this a phenomenological model in the nonlinear
analysis software OpenSees (McKenna et al. 2010). This new model is designated as the
HyperbolicQy model. The nonlinear SSI phenomenological model was added to the nonlinear
finite element model of the building. Then, the ground motion duration effects are studied on the
SSI model of the SMRF buildings by subjecting them to spectrally equivalent short- and long-
duration ground motions sets.

5.3 Model Development

5.3.1 Soil Models


In this study, the buildings are assumed to be supported by a layer of clay loaded seismically
under undrained conditions. The variability of the undrained shear strength su, at each footing
location and between footings, is considered, and affected by a random fluctuating component
that characterizes the spatial variation of the soil properties. The basic material parameters can be
defined in terms of an amplitude (deviation from trend), characterized by the coefficient of
variation (COV), and horizontal scale of fluctuation, δh. Phoon (1995) and Phoon and Kulhawy
(1999) presented guidelines to consider the inherent soil variability and, proposed ranges of
values for coefficient of variation and horizontal scale of fluctuation of soil properties. The
following assumptions are adopted for the considered soil-foundation systems in this work: (1)
rigid shallow footings mobilize a “mean” su at failure, acting across the width of the footing and
within the depth of influence (i.e., representative volume of influence of approximately B3); (2)
su is assumed to be constant at each footing location, but the value is random based on su,mean,
COV(su) and scale of fluctuation, δh, over horizontal distance x; (3) the Random Field Theory
(Fenton and Vanmarcke 1990), which uses autocorrelation with a sine function, was considered
to simulate the spatial variability of the undrained shear strength given by:

 
su  x   su ,mean  A  sin  x  c (5.1)
 h 
144

where A corresponds to the amplitude of su, which is equal to the sum of the su,mean with a
random su (with predefined COV), and c represents an uniformly distributed random offset to
ensure that all values of su obtained are random.

The mean undrained shear strength su,mean of 70 kPa was assumed in this study. Figure 5.1
shows an example of the variability of su along a horizontal distance of 50 m (COV of 15% with
δh= 10 m).

To define the bearing pressure model parameters, the soil variability and the uncertainty of
the model are taken into account considering two values of COV (15% and 30%) and two
autocorrelation distances (10 and 30 meters), leading to four combinations of inherent soil
variability. The values of COV were chosen based on the range of values recommended by
Phoon and Kulhawy (1999), and the values of δh used correspond approximately to the length of
one and three bays of the buildings studied, so that the soil variability is reflected among
footings.

100
COV(su) = 15%, δh = 10 m
90
su , kPa

80
70
60
su,mean = 70 kPa
50
0 10 20 30 40 50
Horizontal Distance x, m

Figure 5.1– Example of spatial variability of undrained shear strength (su,mean = 70 kPa,
COV = 15% and δh = 10 m).

5.3.2 Development of the nonlinear SSI model

5.3.2.1 Bearing pressure-displacement model for shallow footings


A serviceability limit state (SLS) model that estimates the bearing pressure-displacement
response of footings supported on fine-grained soil with undrained response was developed in
Huffman et al. (2015). The Huffman et al. (2015) model estimates the bearing pressure versus
undrained displacement based on the results from shallow footing load test results selected from
the database compiled by Strahler and Stuedlein (2014), considering the following conditions:
145

(1) large enough displacements were reached to estimate nonlinear response, (2) tests that
provided sufficient description of load test set up and rigid footing behavior were considered, (3)
accurate undrained shear strength measurements could be obtained in-situ or laboratory tests, and
(4) footing embedment depth was less than 4B’ (B’ is equivalent footing diameter), indicating
“shallow” spread footing behavior. Among the tests considered, the load-deformation curves
used to fit the model come from tests for relatively uniform soils within the depth of footing
influence, and plastic soils that mobilized undrained shear strength. After normalizing the
bearing pressure by a reference capacity and the displacement by the equivalent footing
diameter, the bearing pressure-displacement curves were obtained as shown in Figure 5.2. This
approach reduces significantly the scatter of data, implying a more accurate determination of the
probability of exceeding the allowable bearing pressure at a given displacement (Huffman et al.
2016).
Normalized Bearing Pressure q/qSTC

1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0.000 0.025 0.050 0.075 0.100 0.125 0.150
Normalized Displacement δ/B’
Figure 5.2 – Bearing pressure-displacement curves used to fit the hyperbolic model
developed in Huffman et al. (2015).

A hyperbolic model was chosen to represent the nonlinear backbone relationship of the
normalized bearing pressure-displacement Δv behavior. Its general formulation is given by:
146

 v 
 B' 
qmob    qref (5.2)

k1  k2  v 
 B' 

where qref is used to normalize the resistance. The slope tangent capacity qSTC with a
displacement offset of 0.03B’ obtained from test data was considered as qref (Huffman et al.
2015). The parameters k1 and k2 are based on the joint distribution obtained from fitting
hyperbolic curves to the load test results. The selection of the parameters distributions are based
on Akaike Information Criterion and Bayesian Information Criterion (Bozdogan 1987) tests.
Detailed information on the validation of qref and development of the backbone model can be
found in Huffman et al. (2015).

In addition to the backbone response of the bearing pressure- displacement model in


compression, the tensile behavior and the cyclic behavior were added to the model that was
developed in OpenSees, as displayed in Figure 5.3 (a). The tensile resistance of the springs is
provided by the weight of the soil placed over the footing. In terms of cyclic response, it is
assumed that there is no stiffness degradation, so that the reload stiffness is the same when
compared with the initial stiffness. When reloading, the model behaves linearly until reaching
the model backbone.

For the study cases considered, the footings are embedded with their thickness equal to the
footing depth of embedment, therefore the tensile resistance is neglected for the specific cases
studied in this chapter.

5.3.2.2 Passive soil resistance model of the shallow footings


The passive resistance of the footing mobilized by the soil is added to the nonlinear SSI
model through an hyperbolic model developed by Duncan and Mokwa (2001). The hyperbolic
load-deflection curve shown in Figure 5.3 (b) is defined by:

 hPR
F PR  (5.3)
1  hPR
 R f PR
K max Fult
147

where FPR represents the mobilized resistance, FultPR the ultimate passive resistance,  hPR the

horizontal footing displacement, Rf is the failure ratio, and Kmax represents the initial stiffness of
the soil. The initial stiffness is calculated based on a solution for loading rectangular area in
elastic half-space proposed by Douglas and Davis (1964), as recommended by Duncan and
Mokwa (2001). The calculation of the ultimate passive resistance is based on the Rankine theory
2
and includes 3-D effects by adding resistance provided by 2su D f to the plane strain resistance,

where Df corresponds to the footing embedment. The failure ratio Rf is defined as the ratio
between the ultimate passive pressure load and the hyperbolic asymptotic value of passive
resistance. Rf can be obtained when measuring the response, or estimated based on experience.
Duncan and Chang (1970) found that values of Rf ranging from 0.75 to 0.95 are appropriate for
hyperbolic representations of stress-strain curves, therefore a failure ratio of 0.85 is assumed. No
initial gap is considered.
The passive resistance is mobilized by the horizontal component of the forces compressing the
soil, not providing any resistance when in tension. Therefore, nonlinear horizontal springs need
to be placed on both sides of the footing.

5.3.2.3 Friction resistance model


The frictional resistance provided by the soil along the base of the footing is added to the
model through a horizontal zero-length element placed at the middle of the footing. A hyperbolic
relationship is assumed to compute the frictional resistance.

 hFr
F Fr  (5.4)
1  Fr
 hFr
Gs Afoot Fult

where Gs is the shear stiffness of the soil, Afoot corresponds to the footing base area, and the
maximum force FultFr is given by:

FultFr  Afoot  su   (5.5)

where α is the friction coefficient. The value assumed for Gs is 8,000 kPa, following the relation
identified by Strahler and Stuedlein (2013) between Gs and the overconsolidation ratio (OCR), in
148

which a value of 8.0 was assumed. A friction coefficient of 0.6, which is commonly used in
practice, was adopted (Perloff and Baron 1976). In OpenSees, the model proposed by
Raychowdhury (2008), and employed later by Harden and Hutchinson (2009), named Tzsimple2
was used after modifying a shape parameter that affects the plastic component of the model
through a trial and error approach until a match between the hyperbolic model defined in
equation (5.4) and the adapted model was reached. In addition, the dashpot coefficient in the
TzSimple2 model was set to zero. The new friction model is hence shown in Figure 5.3 (c).

Mobilized bearing Friction resisting force FFr


pressure qmob Passive resisting
force FPR

gap

Displacement  h
Fr
Displacement v Displacement  h
PR

(a) (b) (c)

Figure 5.3–(a) Bearing pressure-displacement foundation model implemented in OpenSees


as the HyperbolicQy material model; (b) HypergolicGap model used to mobilize the passive
resistance of the footing (Duncan and Mokwa 2001); and (c) Modified TzSimple2 model to
capture the frictional component of load and displacement along base of the footing.

5.3.2.4 OpenSees implementation


The bearing pressure-displacement model described above and shown in Figure 5.3 (a) was
implemented in OpenSees as a new HyperbolicQy material model. The code for this
implemented material model is shown in Appendix.

5.4 Building Models

The SMRF buildings studied in Barbosa et al. (2017) are a 3-story and a 9-story that were
designed for Seattle, WA, USA, as part of the SAC steel project (Venture 2000) using pre-
Northridge codes (UBC 1997). The external frames of the 3-story and 9-story buildings were
designed to resist the lateral (seismic) loads, and the interior frames to resist the gravity loads. A
detailed description of the buildings is available in Gupta and Krawinkler (1999).
149

In the interest of brevity, only crucial information of the nonlinear finite element models is
provided herein. For complete details on the models, the reader is referred to Barbosa et al.
(2017).

Two-dimensional centerline nonlinear finite element models of one external frame of each
of the buildings were developed in OpenSees. Floor masses were applied to beam-column joints,
whereas loads were applied to beam-columns joints and beam spans. Rayleigh damping was
assigned to the models considering a damping ratio of 2% at the fundamental period T1 of the
intact structure of each building and at T = 0.2 sec. The columns were modeled using nonlinear
force-based fiber-section beam-column elements, in which an elasto-plastic constitutive law with
a 3% kinematic hardening was assigned to each fiber. Beams were modeled through finite length
plastic-hinge beam-column element (Scott and Fenves 2006; Ribeiro et al. 2015). A modified
version of the model labeled FMRB (finite-length modified Gauss–Radau with bilinear model),
also known as the modified Ibarra–Medina–Krawinkler (ModIMK) deterioration model (Ibarra
and Krawinkler 2005; Lignos and Krawinkler 2011) is employed to model the plastic hinge
behavior. The original model was validated by Barbosa et al. (2017) through comparison of
eigenvalues and nonlinear static pushover analyses results obtained in OpenSees with the
FEMA355 M1 models. Figure 5.4 shows a scheme of the buildings studied including two
member sections, and Figure 5.5 the general SSI model divided by its components explained in
this section.
150

24.4
Story Column Ext. Column Int. Beam
43.9
9 W24x131 W24x131 W24x62
8 W24x131 W24x162 W24x62
660.4
7 W24x162 W24x162 W24x76
6 W24x162 W24x207 W24x76
5 W24x207 W24x207 W27x94 332.7
4 W24x207 W24x299 W27x94 [mm]
3 W24x229 W24x229 W30x108 W24x229
2 W24x229 W24x229 W30x116
17.3
1 W24x229 W24x229 W30x108
B1 W24x229 W24x229 W30x108

11.2

607.1
Story Column Ext. Column Int. Beam
3 W14x159 W14x176 W18x40
2 W14x159 W14x176 W24x84
228.3
1 W14x159 W14x176 W24x76 [mm]
W24x76

(a) (b) (c)

Figure 5.4 – (a) Two-dimensional models of steel moment resisting frames buildings; (b)
Identification of all column and beams sections; (c) Column (fiber section) and beam (linear
section) sections.

Friction resistance model


Modified TzSimple2 material

Passive resistance model


Passive resistance model HyperbolicGap material
HyperbolicGap material
Bearing pressure model
HyperbolicQy material

Figure 5.5 - General model of the SSI model implemented.

5.5 Number of springs to assign to the bearing pressure model

When defining the number of vertical HyperbolicQy springs to be used, a balance between
reasonable time of analysis and accuracy of results is sought. The long-duration ground motion
from Pair #6 (Hualele) of Maule, Chile 2010 earthquake, was applied to a reference model of
151

each building. The SSI model added to the structural model of the 3-story building considers five
different numbers of springs: 7, 21, 35, 51, and 101 vertical springs with equal tributary area
where the bearing pressure-displacement model is assigned. Only four different numbers of
springs were considered for the 9-story building model, including the 21, 35, 51, and 101
springs. The soil-structure systems were subjected to the exact same ground motion and local
and global structural responses were evaluated for each model. The number of springs after
which the results start converging is selected to use for all analyses.

Figure 5.6 shows the deviations of the results obtained regarding the 3-story building in
terms of (a) angular distortion, (b) interstory drift ratios, and (c) the moment at the end of the left
bay beam for all floor levels. The results are normalized to the results obtained when considering
the 101 springs. From observation of these results, it can be seen that for the 3-story building,
convergence in the results is obtained when considering 35 springs, and hence 35 springs were
considered for all seismic analyses of the 3-story building. For the 9-story building, Figure 5.7
shows negligible deviations were obtained for all trials independently of the number of springs
used. Therefore 21 springs were used to assign the bearing pressure-displacement behavior to the
9-story building footings since it corresponded to the smallest number of springs considered.
This may be related to the fact that the horizontal displacements were restrained at the first floor
level. Modeling of the soil-retaining wall effects should be considered in future works.
152

Angular distortion envelope


20%
Deviation Bay 4 - 5
Bay 3 - 4
0%
Bay 2 - 3
-20% (a) Bay 1 - 2

Interstory drift envelope


12%
Deviation

0%

-12% (b)

Maximum beam moments


12%
Deviation

3rd story
0%
2nd story
-12% (c) 1st story
7 21 35 51
Number of vertical springs

Figure 5.6 - Deviations obtained (from analysis with 101 springs in each footing) for the
calibration of the number of springs to use at each footing of the 3-story building: (a) Angular
distortion, (b) Interstory drift ratio, (c) Peak beam end-moments.
153

Angular distortion envelope Bay 5 - 6


0.040% Bay 4 - 5
Deviation

Bay 3 - 4
0.000% Bay 2 - 3
(a) Bay 1 - 2
-0.040%

Interstory drift envelope

0.040% 9th story


Deviation

8th story
0.000% (b)
7th story
-0.040% 6th story
5th story
Maximum beam moments 4th story
0.040%
3rd story
Deviation

2nd story
0.000% (c)
1st story
-0.040%
21 35 51
Number of vertical springs

Figure 5.7 – Deviations obtained (from analysis with 101 springs in each footing) for the
calibration of the number of springs to use at each footing of the 9-story building: (a) Angular
distortion, (b) Interstory drift ratio, (c) Peak beam end-moments

5.6 Methodology

5.6.1 Foundation Design


For the footing design, a bearing pressure on the footing equal to the allowable bearing
pressure was assumed, considering a factor of safety FSv equal to 2.5, represented by the
following equation:

Q 1
FSv 2
 su  Nc  Fcs  Fcd  q  N q  Fqs  Fqd  B    N  F s  F d (5.6)
B 2

where B represents the footing width (assumed square), and Q corresponds to the vertical load
applied to the footing. On the right hand side of the equation, defining the allowable bearing
pressure, q is the effective stress at the base of the footing, γ = 19.7 kN/m3 is the unit weight of
154

the soil, Nc, Nq, and Nγ correspond to the bearing capacity factors proposed by Meyerhof (1963),
Fcs, Fqs, and Fγs are shape factors also from Meyerhof (1963), and lastly Fcd, Fqd, and Fγd are
depth factors suggested by Hansen (1970). Table 5.1 presents dimensions considered for the
designed footings.

Table 5.1 – Footing dimensions of buildings studied


3-story building 9-story building
Edge Middle Edge Middle
columns columns columns columns
Dimensions B=L (m) 1.68 2.15 2.55 3.25
Thickness H (m) 0.60 1.35

5.6.2 Nonlinear SSI model parameter generation


Monte Carlo simulations are performed to generate the necessary bearing pressure model
parameters for the system soil-foundation. To model the uncertainty of the model parameters
together with the spatial variability of the soil parameters, for each of the four combinations of
COV and δh, 50,000 sets of parameters that include k1, k2, qSTC, ultimate passive resistance FultPR

in both sides of the footing, and su, are generated for each footing.

5.6.2.1 Parameter dependence and copulas


The parameters k1, k2, and qSTC are correlated to each other as shown in Huffman et al.
(2015). Therefore, copula functions were used to account for the multivariate dependence when
applying the Monte Carlo simulations. Dependence between three or more variables can be
accounted for using a multivariate copula, or a string of bivariate copulas, known as vine copula
(Joe 1996; Aas et al. 2009; Brechmann et al. 2013). The vine copula approach was selected here
to represent the dependence between k1, k2, and qSTC, providing a better representation of the
correlation between variables (Brechmann et al. 2013). All considerations on the selection of the
bivariate copulas included in the study, and a summary of the selected copula types, is provided
by Huffman et al. (2015).

5.6.2.2 Sampling methodology


Each seismic analysis considers one ground motion with a determined intensity level, and
one set of footing parameters to define the SSI scenario, which are characterized by a specific set
155

that includes k1, k2, qSTC for the bearing pressure model, FultPR at left and right sides of the footing

for the passive resistance model and the su necessary for the friction resistance model. Since
uncertainty in these parameters is explicitly accounted for, a large number of SSI scenarios need
to be considered. However, each analysis has a high computational cost due to the model
complexity and small time-step used to obtain accurate time-history response. Consequently, a
Latin Hypercube Sampling methodology (McKay et al. 1979) was selected to reduce the number
of analyses and provide a balance between computational cost and good characterization of the
soil uncertainty and its effect in the structural response. Thus, for each combination of COV and
δh, 50,000 SSI scenarios (starting data) were used to initially characterize the soil uncertainty.
Then, a small sample of SSI scenarios (reduced data) was picked from the starting data. The
definition of the sample size of the reduced data needs to consider the possibility of keeping the
soil variability characterized by the starting data, and the feasibility of the analyses that are
generated. Nonetheless, the variability imposed by the final sample of SSI scenarios selected
should be similar when compared with the variability characterized by the starting sample.

To obtain a reduced data set for analysis, the following step-by-step procedure was applied:

Step 1: Generate the starting data: 50,000 SSI scenarios;

Step 2: Compute the starting data correlation matrix between the various parameters. Note
that the correlations k1-k2 and k1-qSTC are nonlinear. In this case, one of the variables was
transformed in order to obtain as close to linear correlation as possible. Therefore, correlations

k1  1 1 q
k2 and k1 STC were considered instead.

Step 3: Generate 10,000 different sets of reduced data of SSI scenarios using the “sample”
function in R-studio (RStudio Team 2015). Each set of reduced data consisted in 20 SSI
scenarios.

Step 4: For each set of reduced data, compute of the correlation matrix.

Step 5: Compute the weighted difference Dwij between the entries of the correlation matrix
associated with the respective reduced data and the starting data correlation matrix
156

Dwij  Wij start  reduced ij where the term Wij in this expression assigns different weights to

entries of matrix that compute the difference between correlations. The sum of the weights Wij
added are equal to 1.0. Seven different ranks of importance were defined ranging from 0% to
2.8%. More details of the analysis can be provided upon request. In this case the highest weight
was assigned to the difference of correlations between parameters used to build the copulas
described in section 5.6.2.1.

Step 6: Select the reduced data that provides the smallest sum of weighted absolute values

of the difference obtained in the previous step: min  Dwij .

Figure 5.8 (a), (b) and (c) show an example of the copulas k1-k2, k1-qSTC, and k2-qSTC, which
were considered for one of the 3-story building footings. The 50,000 original data points are
shown as grey circles and the 20 sets selected after the reduction procedure was applied, shown
as red circles. Figure 5.8 (d) and (e) plot the correlations referred in Step 2 of this section,

k1  1 and 1  qSTC where the variables were transformed. The example in this figure was
k2 k1
developed considering the soil spatial variability and model uncertainty defined with COV =
15% and δh = 10 m. Similar plots and results could be shown for the other footings (see
Appendix for additional parameter sets selected).
157

(a) (b) (c)

(d) (e)
Figure 5.8 - Copulas showing the parameters correlations used for footing #1 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC

5.6.3 Structural Analysis


For each SSI scenario of the reduced dataset, a multiple step structural analysis procedure
was implemented. It is worth noting that depending on the type of analysis, different gravity
loads were applied following the guidelines of ASCE7-10 and ASCE 41-13 (ASCE 1994, 2014).
The steps include:

Step 1: A serviceability analysis is performed; therefore, both dead and live loads are
applied to the buildings. Settlements are measured at each footing level, and angular
distortions (differential settlements over the bay length) are computed and evaluated.
158

Step 2: Before applying the seismic load, only the dead loads are applied to the
buildings, and an eigenvalue analysis is performed to estimate modal periods. The variability
of the periods of the soil-structure system is expected to increase as compared to the fixed-
base model.

Step 3: The buildings are loaded seismically, subjected to the ground motions shown in
Table A9 without considering any ground motion scaling. The first seismic analysis
performed in this step evaluates the statistical importance of ground motion duration on the
response measures (e.g. peak deformations) and damage indices. However, the results
obtained in this first analysis are dependent not only on the significant duration, D5–75, but
also on ground motion intensity, Sa(T1).

Step 4: Incremental dynamic analyses (IDA) are then performed, in which each
individual soil-structural model developed using the reduced parameter dataset is subjected
to 10 short- and 10 long-duration ground motions, at least at 20 identical intensities. The
ground motions selected are underlined in Table A9. For each intensity considered in the
IDA, the effect of ground motion duration is isolated. To do so, the short-duration ground
motions are first scaled so that the response spectrum shape of each short-duration motion
matches the response spectrum of a long-duration ground motion, where “response spectrum
shape matching” here refers to the selection of a scale factor that minimizes a root mean
square of the difference between the natural algorithm of the spectral values (Chandramohan
et al. 2016; Barbosa et al. 2017). Each of the motions of the matching pair is then scaled
identically for the development of the IDA curves, in which the evolution of the measured
response is evaluated as the seismic intensity increases. For comparison between the fixed-
base building models and the building models in which the SSI is incorporated, no period
range of interest is considered, matching both short- and long-duration ground motions
response spectrum across the whole spectrum. The non-consideration of a structural period
range of interest is due to shift in the fundamental period of the building model with SSI.

For steps 3 and 4 described in this section, the response measures recorded are the peak
interstory drift ratios (IDRs), residual interstory drift ratios (RIDRs), energy dissipated, and
number of inelastic excursions at every beam section. The spectral acceleration at the
159

fundamental period of the fixed-base intact structure, Sa(T1Fixed-base), is taken as the intensity
measure for all cases, in order to facilitate the comparisons across models. Structural damage is
computed based on the performance of beams only, since they provide the largest contribution to
the damage, since the buildings were designed based on a strong column-weak beam concept,
and little to no damage is expected in the relatively stocky columns (Newel and Vang, 2008;
Ribeiro et al. 2014). The element damage index is taken as the maximum of the section damage
index of both hinges in a beam element, and the damage index of the global structure is defined
as the average of the element’s damage indices weighted by their hysteretic energy dissipated, as
explained in detail in Barbosa et al. (2017).

5.7 Results and discussion

5.7.1 Modal Analysis


When considering the SSI effects, an increase in the building fundamental periods of
vibration are expected when compared to the fundamental periods of vibration of the fixed-base
models (e.g Raychowdhury, 2011). The three first periods (T1, T2, T3) of both 3-story and 9-story
buildings are shown in Figure 5.9. The results are plotted as a ratio of the fundamental period of
the fixed-base model after gravity load is applied, in which the vertical lines represent the fixed-
base periods. The continuous lines correspond to the periods of the structure after application of
the gravity loads. In this figure, each marker identifies one result from an individual analysis for
a single parameter set considered, and therefore the plots for each marker type represent the
likelihood (Benjamin and Cornell 1970) of the periods of the structural model when SSI effects
are considered. The disparity between the periods of the fixed-base buildings and the mean
periods of the buildings with SSI can be clearly observed. From observation of the results
displayed in the figure, it can be inferred that the changes in the periods are due to application of
gravity loads and/or when considering SSI effects.

From close inspection of the values of the periods of vibration regarding the 3-story
building, the fundamental period of the fixed-base model T1Fixed-base increases from 1.33 to 1.36
sec when applying the gravity loads alone. When looking to the eigenvalues associated with the
use of the SSI models, the mean values of T1 are approximately 2.00 and 2.20 sec before and
after applying the dead load, respectively. When looking at the fixed-base model, the second
160

mode period T2Fixed-base is equal to 0.43 sec and the third mode period T3Fixed-base is equal to 0.22
sec. It is interesting to note that there are negligible differences between the eigenvalues before
and after the gravity load is applied. However, when considering SSI effects, the mean value of
T2 is 0.65 sec and T3 is 0.58 sec before applying the dead load, and T2 is 0.95 sec and T3 is 0.75
sec after applying the gravity loads.

T1 Fixed-base
3-Story building
Likelihood

T2 Fixed-base
T3 Fixed-base
T1 Fixed-base after gravity loads
T2 Fixed-base after gravity loads
T3 Fixed-base after gravity loads
9-Story building
T1 Soil-structure
Likelihood

T2 Soil-structure
T3 Soil-structure
T1 Soil-structure after gravity loads
T2 Soil-structure after gravity loads
0.0 0.5 1.0 1.5 2.0 2.5 3.0
T3 Soil-structure after gravity loads
Ti / [ T1Fixed-base(after gravity load) ]

Figure 5.9 – First three periods of vibration of the building models with different scenarios
normalized by the fundamental period of the fixed-base structure, and the SSI models of the: (a)
3-story building, (b) 9-story building.

From eigenvalue analysis on the 9-story building, T1Fixed-base is 2.69 sec increasing to 2.79
sec after the gravity load is applied. With SSI effects, T1 mean values are 2.95 sec and 3.85( sec
before and after the gravity loads are applied, respectively. In terms of higher modea)periods,
T2Fixed-base is 1.00 sec and T3Fixed-base is 0.55 sec, which remain essentially unchanged after the
application of the gravity load. T2 is not affected by the SSI effects before the building is loaded,
but after the gravity load, T2 is equal to 1.50 sec. The mean values of T3 of the SSI model( are
0.80 sec and 1.05 sec before and after the gravity load, respectively. In summary, b)
it can be
concluded that the parameters associated with the spatial variability and model uncertainty that
are assigned to each SSI scenario induce the differences in the results obtained from modal
analysis.
161

5.7.2 Serviceability analysis


Figure 5.10 and Figure 5.11 show the cumulative distribution functions (CDF) of the angular
distortions obtained at the foundation level for the 3-story and 9-story buildings, respectively.
The ordinate values indicate the probability of not reaching the angular distortion displayed in
the abscissa axis, i.e. the percentage of analyses in which smaller values were obtained.

When analyzing these results, there are some limit values of angular distortion that should
be taken into account (Skempton and MacDonald 1956; Stuedlein and Bong 2017): (1) 1/500 is
considered as the magnitude of angular distortion that is generally associated with no observed
damage, and it has become identified as the “allowable” magnitude of angular distortion; (2) an
angular distortion of 1/300 has been defined as the limit that characterizes exceedance of the
serviceability limit state (SLS); and (3) 1/150 defines the value that corresponds to the ultimate
limit state (ULS), associated with onset of structural damage.
"allowable"

ULS

"allowable"

SLS
SLS

ULS
CDF (Angular Distortion)

CDF (Angular Distortion)

1.0 1.0
0.8 0.8
0.6 0.6

0.4 0.4 COV 15%&&δhδv


COV == 15% = 10 mm
= 10
Bay 1 - 2
Bay 2 - 3 COV 15%&&δhδv
COV == 15% = 30 mm
= 30
0.2 Bay 3 - 4 0.2 = 30% & δ
COV = 30% & hδv = 10
COV = 10 mm
Bay 4 - 5 COV 30%&&δhδv
COV == 30% = 30 mm
= 30
0.0 0.0
0 0.002 0.004 0.006 0.008 0.01 0 0.002 0.004 0.006 0.008 0.01
Angular Distortion Angular Distortion

(a) (b)

Figure 5.10 - Cumulative density functions (CDF) of the angular distortion values obtained
for the 3-story building, considering the “allowable”, service and ultimate limit states: (a) each
bay, considering all variability cases, (b) each variability type for the left bay
162

ULS
SLS
"allowable"
"allowable"

SLS

ULS
1.0

CDF (Angular Distortion)


1.0
CDF (Angular Distortion)

0.8
0.8
0.6
0.6 Bay 1 - 2
Bay 2 - 3
0.4 Bay 3 - 4
0.4 COV 15%& &δhδv
COV==15% = 10
= 10 m m
Bay 4 - 5
COV 15%& &δhδv
COV==15% = 30
= 30 m m
0.2 Bay 5 - 6
0.2 COV 30%& &δhδv
COV==30% = 10 m
= 10 m
COV 30%& &δhδv
COV==30% = 30 m
= 30 m
0.0 0.0
0 0.002 0.004 0.006 0.008 0.01 0 0.002 0.004 0.006 0.008 0.01
Angular Distortion Angular Distortion

(a) (b)

Figure 5.11 - Cumulative density functions (CDF) of the angular distortion values obtained
for the 9-story building considering the “allowable”, service and ultimate limit states: (a) each
bay, considering all variability cases, (b) each variability type for the left bay

From observation of results shown in Figure 5.10 (a) and Figure 5.11 (a), one can see that the
probability of not reaching the “allowable” angular distortion (1/500) is 75% for the 3-story
building and 50% for the 9-story building. When looking to the other limit states according to
Skempton and MacDonald (1956) (SLS – 1/300, and ULS – 1/150), it can be seen that the
probability of the 3-story building not reaching substantial damage is more than 90% and pretty
close to 100%, respectively. For the 9-story building, the results displayed in Figure 5.11 (a)
show around 70% mean probability of not reaching substantial damage (the SLS) and more than
90% of not reaching the ultimate limit state. Note that these are obtained for two levels of COV
and scale of fluctuation, and therefore results are indicative of a adequate design of the
foundations. When analyzing the cumulative distribution function (CDF) for the four
combinations of spatial variability and model uncertainty (different COV or/and δh) displayed on
Figure 5.11 (b), it can be seen that, for the 9-story building, greater COV tends to increase the
median angular distortions of the building, and greater scales of fluctuation also lead to higher
angular distortion, while for the 3-story building differences are negligible when changing COV,
as shown in Figure 5.10 (b). The reason for this observation might be due to the fact that the 9-
story building has greater periods, and when increasing the coefficient of variation on the soil
parameters computations, very critical cases can be obtained reflected by the structural periods.
In Figure 5.9 can be seen that there are several cases where the fundamental period increases
163

more than 50% resulting in fundamental periods greater than 4 sec, and is expectable that
systems with these modal characteristics do not verify the limits here defined. However, to better
understand the sources for these differences, this topic should be explored in great detail in the
future.

5.7.3 Preliminary results and damage assessment calibration


Peak IDR and global damage indices were obtained for each ground motion and SSI
scenario, from combination of the 44 ground motions with the 80 different SSI scenarios, due to
variability imposed in the SSI model, a total of 2560 different responses are obtained in terms of
peak IDR, DIP&A and DIR&V, when SSI effects are considered. The fixed-base models are equally
subjected to all 44 ground motions.

From the results obtained, ratios of the mean peak IDR or DIP&A from analyses with SSI
models relative to the peak IDR/DIP&A obtained for the fixed-base model are computed for each
SSI scenario (peakIDRSSImean/peakIDRFixed-base, and DIP&ASSImean/DIP&AFixed-base), and are related
with the ratio of fundamental periods that represent the increase of the fundamental period of
vibration when considering SSI (T1SSI/T1fixed-base). Distinct results are obtained between both
buildings studied. For the 3-story building, an increase of deformation demands when
considering SSI effects is evident when subjected by either short- or long-duration ground
motions. On the other hand, the damage index tends to be reduced when considering SSI effects,
mostly when subjected by long-duration ground motions. These observations agree with results
obtained in other works regarding low rise buildings (Moghaddasi, 2010; but are in contrast to
Raychowdhury, 2011). For the 9-story building, similar results are observed regarding the peak
interstory drifts. An evident increase of mean peak interstory drifts is observed as the
fundamental period of the structure increases. In contrast with the results observed for the 3-story
building, the global damage represented by means of DIP&A, is higher when compared with the
fixed base model. It is worth noting that the mean damage indices and the peak IDR obtained for
the 9-story building are more noticeable when the buildings are subjected to short-duration
ground motions.

In a second phase, the results obtained are represented as 3-D surfaces that relate the ground
motion significant duration D5-75, the spectral acceleration at the fundamental period of the fixed-
164

base model Sa(T1Fixed-base), versus peak IDR, DIP&A, and DIR&V as shown in Figure 5.13 and
Figure 5.14 for the 3-story and 9-story buildings, respectively. In Figure 5.13 can be seen that the
effect of duration captured when ground motions are applied to fixed-base models, is also
captured when considering SSI effects on the 3-story building. The 3-D surfaces follow similar
trends, showing a clear tendency to higher demands at longer durations in terms of peak IDR and
DIP&A when looking to low seismic intensities. On the other hand, DIR&V are greater for greater
intensities, when subjected to long duration motions, even though not dealing with significant
levels of damage. From observation of these results, the level of structural damage tends to
reduce when incorporating SSI effects, while peak IDRs tend to increase.

The results relative to the 9-story building, displayed in Figure 5.14, show an evident effect
of ground motion duration in both fixed-base and SSI cases, showing similar surfaces. However,
the ground motion duration have greater impact when SSI effects are considered, shown by
steeper surfaces observed in which greater peak IDR and damage indices are obtained for greater
durations observed for greater seismic intensities. The SSI effects in this building are also
responsible for greater peak IDRs and damage indices when compared with the fixed-base
models. The approach followed in the modeling of the basement, which is part of the 9-story
building, has influence on the damage estimated. Since the building is horizontally restrained at
the 1st floor level, the effects of SSI may not be properly captured along the horizontal direction.
The increase of damage observed might be explained by the additional force demands due to the
horizontal fixity.
165

peakIDRSSI/peakIDRFixed-base 1.6 1.6


3-Story building

DIP&ASSI/DIP&AFixed-base
1.2
1.2
0.8

0.8 0.4
1.4 1.6 1.8 1.4 1.6 1.8
T1 /T1Fixed-base
SSI T1SSI/T1Fixed-base

(a) (b)
2.5 2.5
peakIDRSSI/peakIDRFixed-base

9-Story building
2.0 DIP&ASSI/DIP&AFixed-base 2.0

1.5 1.5

1.0 1.0

0.5 0.5
1.0 1.5 2.0 1.0 1.5 2.0
T1SSI/T1Fixed-base T1SSI/T1Fixed-base

(c) (d)
Long-duration Short-duration
Linear (Long-duration) Linear (Short-duration)

Figure 5.12 – Variation of the deformation demands and global damage on the buildings
with increase of fundamental period of the soil-structure system: (a) 3-story building: ratio of
mean peak IDR considering SSI with fixed-base model peak IDR; (b) 3-story building: ratio of
mean DIP&A considering SSI with fixed-base model DIP&A; (c) 9-story building: ratio of mean
peak IDR considering SSI with fixed-base model peak IDR; (d) 9-story building: ratio of mean
DIP&A considering SSI with fixed-base model DIP&A;
166

(a) (b) (c)

(d) (e) (f)


Figure 5.13 – 3-story building responses and surface response functions of the significant
duration D5-75 and spectral acceleration at the fundamental period of the fixed-base model
Sa(T1Fixed-base) for the 44 unscaled ground motion and the 80 SSI different cases (when
considered): (a) Fixed-base model global DIP&A, (b) Fixed-base model global DIR&V, (c) Fixed-
base model peak IDR, (d) SSI model global DIP&A, (e) SSI model global DIR&V, (f) SSI model
peak IDR
167

0.6 5
0.6 0.5 5 4
0.6

Peak IDR, %
0.4 0.4 4
DIP&A

0.4 3 3

DIR&V
0.2 0.3 2
0.2 2
0.2 1
0
0 0
80 100 0.1 1
10-1 80 100 80 100
40 10-1
10-2 40 40 10-1
0 10-3 10-2 0 10-2
0 10-3 0 10-3

(a) (b) (c)


5
0.6
0.6 0.6 0.5 5 4

Peak IDR, %
0.4 4
0.4
DIP&A

0.4
DIR&V

3 3
0.2 0.3 2
0.2
0 0.2 1 2
0
80 100 0
40 10-1
80 100 0.1 80 100 1
10-2 40 10-1 0 40 10-1
0 10-3 0 10-3 10-2 10-2
0 10-3

(d) (e) (f)


Figure 5.14 - 9-story building response function of the significant duration D5-75 and spectral
acceleration at the fundamental period of the fixed-base model Sa(T1Fixed-base) for the 44 unscaled
ground motion and the 80 SSI different cases (when considered): (a) Fixed-base model global
DIP&A , (b) Fixed-base model global DIR&V, (c) Fixed-base model peak IDR, (d) SSI model
global DIP&A, (e) SSI model global DIR&V, (f) SSI model peak IDR.

5.7.4 Influence of SSI on duration effects

5.7.4.1 Deformation and energy demands


Time-histories of IDR and energy dissipated by yielding at one beam end of both 3- and 9-
story buildings are analyzed in this subsection. A long-duration (Maule, Chile 2010 recorded in
Hualane) and its respective scaled short-duration (Santa Barbara 1978, recorded in Cachuma
Dam Toe) ground motions are applied to both buildings under fixed-base conditions, and
considering SSI effects (one scenario).

The results obtained from this analysis aim to understand how SSI affects the localized
response of determined regions of the buildings, and consequently the global structural behavior
that leads to the damage and deformation observed. Figure 5.15 (a) shows the time-history plots
168

in terms of IDR and energy dissipated at the beam end of the 1st bay of 3rd and 1st floors for the
3-story building. Figure 5.15 (b) shows the same time-history plots for the 8th (1st above the
basement) and 2nd floors of the 9-story building.
169
4%
3rd Story
IDR
0%
Fixed-base model Fixed-base model
-4%
0 25 50 0 50 100 150 200
time t, sec time t, sec
4%
IDR

0%
SSI model SSI model
-4%
0 25 50 0 50 100 150 200
time t, sec time t, sec
4%
1st Story
IDR

0%
Fixed-base model Fixed-base model
-4%
0 25 50 0 50 100 150 200
time t, sec time t, sec
4%
SSI model
IDR

0%
SSI model
-4%
0 25 50 0 50 100 150 200
time t, sec time t, sec

(a)
2%
8th Story
IDR

0%
Fixed-base model Fixed-base model
-2%
0 25 50 0 50 100 150 200
time t, sec time t, sec
2%
IDR

0%
SSI model SSI model
-2%
0 25 50 0 50 100 150 200
time t, sec time t, sec
2%
2nd Story
IDR

0%
Fixed-base model Fixed-base model
-2%
0 25 50 0 50 100 150 200
time t, sec time t, sec
2%
IDR

0%
SSI model SSI model
-2%
0 25 50 0 50 100 150 200
time t, sec time t, sec

(b)

Figure 5.15 – History of IDRs during one short-duration and one long-duration ground
motions considered both fixed base and SSI models: (a) 3-story building 1st and 3rd stories; (b) 9-
story building 2nd and 8th stories.
170

1200 1200
3-story 9-story
1000 Fixed-base 1000
building building
SSI: Beams+V.Spr+H.Spr
800 800
Eh, kNm

Eh, kNm
SSI: Beams+V.Spr+H.Spr 600 Fixed-base
600
SSI: V.Spr
400 400
SSI: V.Spr
200 200 SSI: Beams
SSI: H.Spr
SSI: Beams SSI: H.Spr
0 0
0 25 50 75 100 125 150 175 200 0 25 50 75 100 125 150 175 200
time t, sec time t, sec

(a) (b)

Figure 5.16 – Energy dissipated by yielding of the beams not considering SSI effects, energy
dissipated by yielding of beams considering SSI effects, and vertical and horizontal springs that
compose the SSI model: (a) 3-story building; and (b) 9-story building.

From observation of the results shown in Figure 5.15 relative to 3-story building, while the
results of the fixed-base case show greater IDRs at the 3rd story, the results regarding the model
with SSI effects show greater IDRs at the 1st story. On the other hand, the quantity of energy
dissipated by yielding at the beam ends of the building model with SSI effects is much smaller
than the energy dissipated in the fixed-base model. These results help to understand the findings
from Figure 5.13 where greater deformation and less damage were observed for the case where
SSI was taken into account.

The results observed for the 9-story building do not reveal significant differences in terms of
peak IDRs when comparing the fixed-base and the SSI models considered. However, the energy
dissipated by yielding of the beams of the fixed-base model is higher than the energy dissipated
by the SSI model, as observed for the 3-story building.

Analyzing the time-history of IDRs of both buildings, residual interstory drifts are observed
after the seismic action. These residual interstory drifts are, in some cases, of similar magnitude
of the interstory drifts observed before the seismic action, induced by the differential settlements
obtained after the gravity loads are applied. The fact that the 9-story building shows the highest
probabilities of obtaining greater angular distortions (when compared with the 3-story building),
as observed in section 5.7.2, explains the grater residual interstory drifts observed in these plots.
171

The reduction of energy dissipated at the beams is explained by the fact that a great amount
of energy is absorbed by the soil, dissipated by yielding of the zero-length elements that define
the SSI model. Figure 5.16 compares the energy dissipated by yielding of the beams for the
fixed-base model with the energy dissipated by yielding when SSI is considered, when subjected
to the Maule 2010 (Hualele) ground motion. The results regarding the SSI model are
representative of one SSI scenario. If neglecting the energy dissipated by the building columns,
which negligible, when SSI is taken into account, the energy is dissipated by yielding of the
beams and both vertical and horizontal zero-length elements that define de SSI model, while for
the fixed-base case, the energy is dissipated by yielding of the beams only. For the fixed-base
model of the 3-story building, the amount of energy dissipated by the beams is greater than the
energy dissipated by all elements of the model that accounts for SSI (beams, vertical and
horizontal zero-length elements), explaining the reduction of damage when considering SSI. On
the other hand, in the 9-story building, the energy dissipated by the model that considers SSI
effects is greater than the energy dissipated by the beams of the fixed-base model, which
explains the greater damage indices obtained when accounting for SSI. Despite negligible energy
dissipated by the horizontal zero-length elements due to the restriction imposed at the 1st floor
level due to the existence of a basement, there is a large percentage of energy that is dissipated
by the vertical SSI elements.

5.7.4.2 IDA results


Median IDAs curves showing the evolution of the peak IDR and RIDR, with the increase of
seismic intensity, associated with short- and long-duration ground motions are displayed in
Figure 5.17 and Figure 5.18, regarding the 3- and 9-story buildings respectively, where both
fixed-base model and model with SSI effects are considered.
172

2.0 2.0

1.5 1.5
Sa(T1Fixed-base)

Sa(T1Fixed-base)
1.0 1.0

0.5 0.5

0.0 0.0
0.0 5.0 10.0 0.0 2.5 5.0
Peak Interstory drift ratio, % Residual drift ratio, %

(a) (b)
Fixed-Base model - Short-duration Model with SSI - Short-duration
Fixed-Base model - Long-duration Model with SSI - Long-duration

Figure 5.17 – IDA curves comparing median responses of fixed-base and SSI models of the
3-story building subjected to short- and long-duration ground motions in terms of: (a) peak
interstory drift ratios, and (b) Residual interstory drift ratios

0.5 0.5

0.4 0.4
Sa(T1Fixed-base)
Sa(T1Fixed-base)

0.3 0.3

0.2 0.2

0.1 0.1

0.0 0.0
0.0 5.0 10.0 0.0 2.5 5.0
Peak Interstory drift ratio, % Residual drift ratio, %

(a) (b)
Fixed-Base model - Short-duration Model with SSI - Short-duration
Fixed-Base model - Long-duration Model with SSI - Long-duration

Figure 5.18 – IDA curves comparing median responses of fixed-base and SSI models of the
9-story building subjected to short- and long-duration ground motions in terms of: (a) peak
interstory drift ratios, and (b) Residual interstory drift ratios

From observation of Figure 5.17, can be seen that the ground motion duration plays an
important role on the 3-story building for both cases: when SSI effects are considered and for the
fixed-base model. When analyzing the influence of the SSI effects, in terms of deformations,
173

both peak and residual drifts are similar when compared with the fixed-base model results, for
smaller intensities. However, for greater seismic intensities, the SSI model tends to show greater
deformations, for the same intensity level, not only in terms of IDRs but also when looking to
RIDRs.

Observing the curves relative to the 9-story building plotted in Figure 5.18, the duration
effects play an important role either considering SSI effects or simply a fixed-based model.
Greater peak IDR and RIDR are obtained for the same seismic intensity level when subjecting to
long-duration ground motions. When looking to the deformation demands imposed on the
building when SSI effects are considered, demands are higher when comparing with the fixed-
base model.
174

1.0 1.0
0.8 0.8
DIP&AMax.member

0.6 0.6

DIP&A
0.4 0.4
0.2 0.2
0.0 0.0
0.0 0.5 1.0 1.5 0.0 0.5 1.0 1.5
Sa(T1Fixed-Base) Sa(T1Fixed-Base)

(a) (b)

1.0 1.0
0.8
DIR&VMax.member

0.8
0.6 0.6
DIR&V

0.4 0.4
0.2 0.2
0.0 0.0
0.0 0.5 1.0 1.5 0.0 0.5 1.0 1.5
Sa(T1Fixed-Base) Sa(T1Fixed-Base)

(c) (d)
Fixed-Base model - Short-duration Model with SSI - Short-duration
Fixed-Base model - Long-duration Model with SSI - Long-duration

Figure 5.19 - Curves comparing median damage indices of the fixed-base and SSI models of
the 3-story building subjected to short- and long-duration ground (a) DIP&AMax.member, and (b)
global DIP&A, (c) DIR&VMax.member, and (d) global DIR&V
175

1.0 1.0
0.8 0.8
DIP&AMax.member

0.6 0.6

DIP&A
0.4 0.4

0.2 0.2

0.0 0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
Sa(T1Fixed-base)
Sa(T1Fxed-base)

(a) (b)

1.0 1.0
0.8 0.8
DIR&VMax.member

0.6 0.6
DIR&V

0.4 0.4
0.2 0.2
0.0 0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
Sa(T1Fxed-base) Sa(T1Fixed-base)

(c) (d)
Fixed-Base model - Short-duration Model with SSI - Short-duration
Fixed-Base model - Long-duration Model with SSI - Long-duration
Figure 5.20 - Curves comparing median damage indices of the fixed-base and SSI models of
the 9-story building subjected to short- and long-duration ground (a) DIP&AMax.member, and (b)
global DIP&A, (c) DIR&VMax.member, and (d) global DIR&V

Analysis of the damage occurred in the buildings subjected to short- and long-duration
ground motion motions, considering or not SSI effects, is performed by observing curves
displayed in Figure 5.19 and Figure 5.20 that show the evolution of the global DIP&A and DIR&V
with the increase of the seismic intensities, as well as curves referred to maximum DIP&A and
DIR&V at one member of the building, relative to the 3-story and 9-story buildings respectively.

From observation of Figure 5.19, the plots relative to DIP&A and DIR&V of the 3-story
building show that ground motion duration plays a fundamental role in terms of damage, either
when considering the maximum damage observed at one structural member DIP&AMax.member and
176

DIR&VMax.member, or in terms of global damage, considering or not SSI effects. Focusing on the
influence of SSI on the results, the DIP&A curves show that SSI effects tend to reduce the damage
at lower levels of damage, so that, the same level of damage is observed at greater seismic
intensities, but collapse is observed at similar seismic intensities. On the other hand, DIR&V
results indicate that SSI reduce damage at all levels of seismic intensity.

For the 9-story building, considering or not SSI effects, the ground motion duration plays a
fundamental role in the structural behavior of the building, with more evident results than the
ones obtained for 3-story buildings. When assessing SSI effects, not only higher deformations
are observed when compared with fixed-base model, but also the local damage evaluated by
means of DIP&AMax.member and DIR&VMax.member, is greater for the same intensity level as can be
observed in Figure 5.20 (a) and (c). In terms of global damage, no significant differences are
observed in Figure 5.20 (b) and (d), when SSI effects are considered.

5.8 Conclusions

The effect of ground motion duration was evaluated on the structural response of steel
moment resisting frame buildings, considering soil-structure interaction effects and compared
with results obtained considering fixed-base models. Horizontal spatial variability of the soil and
soil model uncertainty were considered on the ultimate passive force in both sides of the
footings, on the parameters that define the bearing capacity model developed by Huffman et al.
(2015), and on the footing friction model. The passive and friction resistance models were
considered in the nonlinear SSI models and implemented in OpenSees. Results that show that:

 A strong impact of the bearing pressure model parameters used on the building periods
was observed. The numerical models that incorporate SSI effects show significantly
higher periods than the fixed-base models.
 Depending on the soil parameters, the gravity load can have a significant effect on the
periods of the soil-structure system.
 Considering soil spatial variability and uncertainty on the SSI model parameters, the
limits of angular distortion that have been considered by the research community and
industry as “allowable” limit, SLS, and ULS on the 9-story building, were not reached for
177

50%, 70%, and 90% of the cases, while the 3-story building verified the same limits for
75%, 90% and 100% of the cases, when considering a FSv = 2.5.
 The energy demands of the buildings reduce considerably when accounting for SSI
effects, as observed by results of the energy dissipated by yielding of the beams.
Considering SSI effects, part of the energy is dissipated by yielding of the the springs that
constitute the SSI model.
 The ground motion duration also plays an important role when SSI effects are considered,
by increasing the estimated deformation demands and damage for longer durations.
 When SSI effects are taken into account, peak interstory drifts and residual interstory
drifts tend to be greater. Depending on the soil parameters considered, some interstory
drift can be observed before the seismic load is applied, induced by high angular
distortions obtained after dead loads are applied.

In terms of future work, an extension of the phenomenological SSI model to 3-D in


order to be applied to 3-D structures should be considered. A greater number of values of
COV and δh should be considered to define a wider soil variability. A different modeling
approach should be considered for the basement present in the 9-story building, where a
different soil-structure interaction problem should be tackled.
178

5.9 References

Aas, K., Czado, C., Frigessi, A., and Bakken, H. (2009). “Pair-copula constructions of multiple
dependence.” Insurance: Mathematics and economics, Elsevier, 44(2), 182–198.

Allotey, N., and El Naggar, M. H. (2003). “Analytical moment-rotation curves for rigid
foundations based on a Winkler model.” Soil Dynamics and Earthquake Engineering, 23(5),
367–381.

Allotey, N., and El Naggar, M. H. (2008). “An investigation into the Winkler modeling of the
cyclic response of rigid footings.” Soil Dynamics and Earthquake Engineering, 28(1), 44–
57.

ASCE. (1994). Minimum design loads for buildings and other structures. American Society of
Civil Engineers, Reston, VA.

ASCE. (2014). “Seismic Evaluation and Retrofit of Existing Buildings: ASCE Standard
ASCE/SEI 41-13.” Reston, VA.

Barbosa, A. R., Ribeiro, F. L. A., and Neves, L. C. (2017). “Influence of earthquake ground-
motion duration on damage estimation: application to steel moment resisting frames.”
Earthquake Engineering and Structural Dynamics, 46(1), 27–49.

Bartlett, P. E. (1976). Foundation rocking on clay soil. Report (University of Auckland, School of
Engineering). Department of Civil Engineering, University of Auckland.

Belejo, A., Barbosa, A. R., and Bento, R. (2017). “Influence of Ground Motion Duration on
Damage Index-based Fragility Assessment of a Plan-Asymmetric Non-Ductile Reinforced
Concrete Builing.” Engineering Structures, 151, 682–703.

Benjamin, J. R., and Cornell, C. A. (1970). “Probability and decision for civil engineers.”
McGraw-Hill, New York.

Boulanger, R. W., Curras, C. J., Kutter, B. L., Wilson, D. W., and Abghari, A. (1999). “Seismic
soil-pile-structure interaction experiments and analyses.” Journal of Geotechnical and
Geoenvironmental Engineering, American Society of Civil Engineers, 125(9), 750–759.

Bozdogan, H. (1987). “Model selection and Akaike’s information criterion (AIC): The general
theory and its analytical extensions.” Psychometrika, Springer, 52(3), 345–370.

Brechmann, E. C., Schepsmeier, U., and others. (2013). “Modeling dependence with C-and D-
vine copulas: The R-package CDVine.” Journal of Statistical Software, 52(3), 1–27.
179

Chandramohan, R., Baker, J. W., and Deierlein, G. G. (2016). “Quantifying the Influence of
Ground Motion Duration on Structural Collapse Capacity Using Spectrally Equivalent
Records.” Earthquake Spectra, 32(2), 927–950.

Chopra, A. K., and Yim, S. C. ‐ S. (1985). “Simplified Earthquake Analysis of Structures with
Foundation Uplift.” Journal of Structural Engineering, 111(4), 906–930.

Cremer, C., Pecker, A., and Davenne, L. (2001). “Cyclic macro-element for soil-structure
interaction: Material and geometrical non-linearities.” International Journal for Numerical
and Analytical Methods in Geomechanics, 25(13), 1257–1284.

Douglas, D. J., and Davis, E. H. (1964). “The movement of buried footings due to moment and
horizontal load and the movement of anchor plates.” Geotechnique, Thomas Telford Ltd,
14(2), 115–132.

Drucker, D. C., and Prager, W. (1952). “Soil mechanics and plastic analysis or limit design.”
Quarterly of applied mathematics, 10(2), 157–165.

Duncan, J. M., and Chang, C.-Y. (1970). “Nonlinear analysis of stress and strain in soils.”
Journal of Soil Mechanics & Foundations Division, 1629–1666.

Duncan, J. M., and Mokwa, R. L. (2001). “Passive earth pressures: theories and tests.” Journal of
Geotechnical and Geoenvironmental Engineering, American Society of Civil Engineers,
127(3), 248–257.

Dutta, A., and Mander, J. B. (2001). “Energy Based Methodology for Ductile Design pf
Concrete Columns.” ASCE Journal of Structural Engineering, 127(12), 1374–1381.

Dutta, S. C., and Roy, R. (2002). “A critical review on idealization and modeling for interaction
among soil--foundation--structure system.” Computers & structures, Elsevier, 80(20),
1579–1594.

Fenton, G. A., and Vanmarcke, E. H. (1990). “Simulation of random fields via local average
subdivision.” Journal of Engineering Mechanics, American Society of Civil Engineers,
116(8), 1733–1749.

Figini, R., Paolucci, R., and Chatzigogos, C. T. (2012). “A macro-element model for non-linear
soil--shallow foundation--structure interaction under seismic loads: theoretical development
and experimental validation on large scale tests.” Earthquake Engineering & Structural
Dynamics, Wiley Online Library, 41(3), 475–493.

Fletcher, D. Q., and Hermann, L. R. (1971). “Elastic foundation representation of continuum.”


Journal of the Engineering Mechanics Division, ASCE, 97(1), 95–107.
180

Gajan, S. (2006). “Physical and numerical modeling of nonlinear cyclic load-deformation


behavior of shallow foundations supporting rocking shear walls.”

Gajan, S., and Kutter, B. L. (2008). “Capacity, Settlement, and Energy Dissipation of Shallow
Footings Subjected to Rocking.” Journal of Geotechnical and Geoenvironmental
Engineering, 134(8), 1129–1141.

Gajan, S., Raychowdhury, P., Hutchinson, T. C., Kutter, B. L., and Stewart, J. P. (2010).
“Application and validation of practical tools for nonlinear soil-foundation interaction
analysis.” Earthquake Spectra, 26(1), 111–129.

Gazetas, G., and Apostolou, M. (2004). “Nonlinear soil--structure interaction: foundation


uplifting and soil yielding.” Proceedings of the 3rd USA--Japan workshop on soil--structure
interaction., Menlo Park, CA.

Gazetas, G., and Mylonakis, G. (2001). “Soil-structure interaction effects on elastic and inelastic
structures.” International Conferences on Recent Advances in Geotechnical Earthquake
Engineering and Soil Dynamics, University of Missouri--Rolla, Rolla, Missouri.

Gupta, A., and Krawinkler, H. (1999). “Seismic Demands for Performance Evaluation of Steel
Moment Resisting Frame Structures.” (132), 1–379.

Hansen, J. B. (1970). A revised and extended formula for bearing capacity. Bull No 28, PP 5-11,
Lyngby, Denmark.

Harden, C. W., and Hutchinson, T. C. (2009). “Beam-on-nonlinear-winkler-foundation modeling


of shallow, rocking-dominated footinǵs.” Earthquake Spectra, 25(2), 277–300.

Huffman, J. C., Martin, J. P., and Stuedlein, A. W. (2016). “Calibration and assessment of
reliability-based serviceability limit state procedures for foundation engineering.” Georisk:
Assessment and Management of Risk for Engineered Systems and Geohazards, 10(August),
1–14.

Huffman, J. C., Strahler, A. W., and Stuedlein, A. W. (2015). “Reliability-based serviceability


limit state design for immediate settlement of spread footings on clay.” Soils and
Foundations, Elsevier, 55(4), 798–812.

Ibarra, L. F., and Krawinkler, H. (2005). “Global Collapse of Frame Structures under Seismic
Excitations.” Evaluation, (152), 1 – 301.

Joe, H. (1996). “Families of m-variate distributions with given margins and m (m-1)/2 bivariate
dependence parameters.” Lecture Notes-Monograph Series, JSTOR, 120–141.
181

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.

Manfredi, G., Polese, M., and Cosenza, E. (2003). “Cumulative demand of the earthquake
ground motions in the near source.” Earthquake Engineering and Structural Dynamics,
32(12), 1853–1865.

McKay, M. D., Beckman, R. J., and Conover, W. J. (1979). “Comparison of three methods for
selecting values of input variables in the analysis of output from a computer code.”
Technometrics, Taylor & Francis, 21(2), 239–245.

McKenna, F., Scott, M. H., and Fenves, G. L. (2010). “Nonlinear Finite-Element Analysis
Software Architecture Using Object Composition.” Journal of Computing in Civil
Engineering, 24(1), 95–107.

Meyerhof, G. G. (1963). “Some recent research on the bearing capacity of foundations.”


Canadian Geotechnical Journal, NRC Research Press, 1(1), 16–26.

Moghaddasi, M., Cubrinovski, M., Chase, J. G., Pampanin, S., and Carr, A. (2011). “Effects of
soil--foundation--structure interaction on seismic structural response via robust Monte Carlo
simulation.” Engineering Structures, Elsevier, 33(4), 1338–1347.

Moghaddasi, M., Cubrinovski, M., Chase, J. G., Pampanin, S., and Carr, A. (2012). “Stochastic
quantification of soil-shallow foundation-structure interaction.” Journal of Earthquake
Engineering, Taylor & Francis, 16(6), 820–850.

Moghaddasi, M., Cubrinovski, M., Pampanin, S., Carr, A. J., and Chase, J. G. (2010). “Soil-
Foundation-Structure Interaction Effects on Nonlinear Seismic Demand of Structures.”
University of Canterbury. Civil and Natural Resources Engineering.

Mylonakis, G., and Gazetas, G. (2000). “Seismic soil-structure interaction: beneficial or


detrimental?” Journal of Earthquake Engineering, World Scientific, 4(03), 277–301.

Nakaki, D. K., and Hart, G. C. (1987). “Uplifting response of structures subjected to earthquake
motions.” EKEH.

Oyarzo-vera, C., and Chouw, N. (2008). “Effect of earthquake duration and sequences of ground
motions on structural responses.” Proceedings of the 10th International Symposium on
Structural Engineering of Young Experts, Changsha, Scottland.

Paolucci, R., Shirato, M., and Yilmaz, M. T. (2008). “Seismic behaviour of shallow foundations:
Shaking table experiments vs numerical modelling.” Earthquake Engineering & Structural
Dynamics, Wiley Online Library, 37(4), 577–595.
182

Park, Y., and Ang, A. (1985). “Mechanistic Seismic Damage Model for Reinforced Concrete.”
ASCE Journal of Structural Engineering, 111(4), 722–739.

Pecker, A., and Chatzigogos, C. T. (2010). “Non linear soil structure interaction: impact on the
seismic response of structures.” Earthquake Engineering in Europe, Springer, 79–103.

Perloff, W. H., and Baron, W. (1976). “Soil mechanics. Principles and applications.”

Phoon, K. (1995). “Reliability-based Design of Foundations for Transmission Line Structures.”


Cornell University.

Phoon, K.-K., and Kulhawy, F. H. (1999). “Characterization of geotechnical variability.”


Canadian Geotechnical Journal, NRC Research Press, 36(4), 612–624.

Raghunandan, M., and Liel, A. B. (2013). “Effect of ground motion duration on earthquake-
induced structural collapse.” Structural Safety, Elsevier, 41, 119–133.

Raghunandan, M., Liel, A. B., and Luco, N. (2015). “Collapse Risk of Buildings in the Pacific
Northwest Region due to Subduction Earthquakes.” Earthquake Spectra, 31(4), 2087–2115.

Raychowdhury, P. (2008). “Nonlinear winkler-based shallow foundation model for performance


assessment of seismically loaded structures.” University of California, San Diego.

Raychowdhury, P. (2011). “Seismic response of low-rise steel moment-resisting frame (SMRF)


buildings incorporating nonlinear soil-structure interaction (SSI).” Engineering Structures,
Elsevier Ltd, 33(3), 958–967.

Reinhorn, A. M., and Valles, R. E. (1995). “Damage evaluation in inelastic response of


structures: a deterministic approach.” National Center for Earthquake Engineering
Research, State University of New York at Buffalo, Buffalo, NY Report No. NCEER-95-xxxx.

Ribeiro, F. L., Barbosa, A. R., Scott, M. H., and Neves, L. C. (2015). “Deterioration Modeling of
Steel Moment Resisting Frames Using Finite-Length Plastic Hinge Force-Based Beam-
Column Elements.” Journal of Structural Engineering, 141(2), 04014112.

RStudio Team. (2015). “RStudio: Integrated Development Environment for R.” Boston, MA.

Ruiz-García, J. (2010). “On the influence of strong-ground motion duration on residual


displacement demands.” Earthquake and Structures, 1(4), 327–344.

Scott, M. H., and Fenves, G. L. (2006). “Plastic Hinge Integration Methods for Force-Based
Beam–Column Elements.” Journal of Structural Engineering, 132(2), 244–252.
183

Skempton, A. W., and MacDonald, D. H. (1956). “The allowable settlements of buildings.”


Proceedings of the Institution of Civil Engineers, Thomas Telford-ICE Virtual Library,
5(6), 727–768.

Song, R., Li, Y., and van de Lindt, J. W. (2014). “Impact of earthquake ground motion
characteristics on collapse risk of post-mainshock buildings considering aftershocks.”
Engineering Structures, Elsevier Ltd, 81, 349–361.

Stewart, J. P., Kim, S., Bielak, J., Dobry, R., and Power, M. S. (2003). “Revisions to soil-
structure interaction procedures in NEHRP design provisions.” Earthquake Spectra, 19(3),
677–696.

Strahler, A. W. (2012). “Bearing capacity and immediate settlement of shallow foundations on


clay.” Oregon State University.

Strahler, A. W., and Stuedlein, A. W. (2013). “Characterization of Model Uncertainty in


Immediate Settlement Calculations for Spread Footings on Clays.” Proceedings, 18th Int.
Conf. Soil Mech. and Geotech. Engrg.

Strahler, A. W., and Stuedlein, A. W. (2014). “Accuracy, Uncertainty, and Reliability of the
Bearing-Capacity Equation for Shallow Foundations on Saturated Clay.” Geo-Congress
2014: Geo-characterization and Modeling for Sustainability, 3262–3273.

Stuedlein, A. W., and Bong, T. (2017). “Effect of Spatial Variability on Static and Liquefaction-
Induced Differential Settlements.” Geo-Risk 2017, 31–51.

Taylor, P. W., Bartlett, P. E., and Wiessing, P. R. (1980). Foundation rocking under earthquake
loading. University of Auckland, Department of Civil Engineering.

Toh, J. C. W., Pender, M. J., and McCully, R. (2011). “Implications of soil variability for
performance based shallow foundation design.” Proc. 9th Pacific Conference on
Earthquake Engineering, Auckland, April, 14–16.

UBC. (1997). “Structural engineering design provisions.” International conference of building


officials.

Venture, S. A. C. J. (2000). “FEMA 355C: State of the art report on system performance of steel
moment frames subject to earthquake ground shaking.” Federal Emergency Management
Agency.
184

6 CONCLUSIONS

6.1 Summary of the Research Performed

Recent long-duration earthquakes such as the ones experienced in the subduction zones
(Chile 2010; Japan 2011; Nepal 2015) have brought the topic of ground motion duration and its
impact on building and bridge response to the research limelight. The effect of long-duration
motions on structural response could be important for regions that have experienced and can
experience again great M8.5+ earthquakes. These regions include, among others, the Pacific
Northwest regions of the United States and Canada, Southern Europe and Northern Africa (in the
boundaries between the African and the Eurasian plate), as well as in Nepal and surrounding
countries (in the boundary of the Indian plate and the Eurasian plate boundaries). Even though
some recent studies have shown that longer duration ground motions may lead to greater collapse
risk, an extensive assessment of the influence of ground motion duration on structural damage is
still lacking in the earthquake engineering literature. As a result, existing seismic risk
methodologies are typically based on response and damage measures calibrated to short-duration
earthquake ground motions.

In the first manuscript presented in this dissertation, the effect of earthquake ground motion
duration was evaluated on the response of a 3-story non-ductile reinforced concrete plan-
asymmetric vintage building. The building model analyzed corresponds to a full-scale building
that was tested at the ELSA laboratories in Europe in January 2004, which was designed to
withstand gravity loads only, therefore not designed to sustain any seismic loading. A nonlinear
finite element model of the 3-story building was developed in OpenSees. The finite element
model was developed specifically with the intent of analyzing the ground motion duration
effects, and validation between experimental results and analytical results presented in this work
showed a reasonable approximation in terms of peak responses as well as number of cycles,
which are important response parameters in this study. The model was also validated in terms of
damage estimated through comparison of the damage observed after the test with damage indices
obtained from numerical analysis. Results from the damage assessment performed to the plan
185

asymmetric 3-D RC building are based on the damage indices by Park and Ang (DIP&A) and by
Reinhorn and Valles (DIR&V), which provided new results for damage assessment on 3-D
buildings models. The main finding from this first study was the important role played by the
ground motion duration in the damage estimation, where the damage estimated was found
greater for longer duration ground motions. These effects are evident for intensities that lead to
collapse of the structure. The ground motion duration did not play an important role on the
displacement and drift responses of this 3-D plan asymmetric RC building.

For the second manuscript presented, the effect of ground motion duration was evaluated on
the response of typical Oregon bridges built between 1950 and 1970, with and without seismic
retrofitting. The bridges analyzed include: (1) the already demolished Mackenzie River Bridge
on Highway Interstate 5, (2) a 3-span regular reinforced concrete bridge with median column
heights, and (3) three irregular bridges with three different column heights, including one column
with the 16th percentile column height, one with the median column height, and one with the 84th
percentile of column height. In this second study, a novel retrofitting solution reported recently
in Lostra (2016) was considered to provide enhanced cyclic performance of the bridge columns.
Thus, since all five bridges analyzed had lap-splices with insufficient length to be able to
adequately transfer forces from the column bars to the footing starter bars. The new retrofit
solution in Lostra (2016) was implemented in OpenSees. Thus, in total, ten bridges were studied.
Models of all non-retrofitted and retrofitted bridges were developed in OpenSees with special
attention dedicated to the modeling of the lap-splice behavior, and on modeling the titanium bars
and the effects of confinement. Cyclic analyses were performed to the column models and
validated with results obtained from laboratory tests of real scale specimens of columns
representative of typical Oregon bridges. Results of the damage assessment performed for all
bridges included peak drift ratios and damage indices by Park & Ang (DIP&A) and Reinhorn &
Valles (DIR&V). The main findings from this second study were: (1) the proposed retrofit solution
applied to the bridges columns provides a significant increase to the structural strength and
deformation capacity; however, the retrofit solution is not as efficient when the bridge has
columns with different lengths along its longitudinal direction versus the ideal case when the
bridge has columns with the same lengths; and (2) the ground motion duration plays an
important role in the damage predicted by the two damage indices in both non-retrofitted and
186

retrofitted bridges, showing greater values of damage index when subjected by long-duration
ground motions. The results shown in this chapter should be considered when revising the
seismic assessment methodologies of as-built and retrofitted bridges by considering the effects of
increased ground motion duration in the bridges structural capacity, and by considering the
bridge irregularity in the design of retrofitting solution to apply to bridge columns.

Lastly, in the third manuscript, the effect of ground motion duration was evaluated on the
structural response of steel moment resisting frame buildings, considering nonlinear soil-
structure interaction effects and compared with results obtained considering fixed-base models.
Horizontal variability of the soil was considered on the ultimate passive force in both sides of the
footings, and on the parameters that define a novel bearing capacity model. The development of
this new bearing model was based on high quality experimental load-bearing pressure tests
(Huffman et al., 2015). In addition, passive and friction resistance models were considered in the
nonlinear SSI model developed, which was implemented in OpenSees. Results indicate that: (1)
the inclusion of nonlinear SSI effects increases the steel moment resisting frame building
measured response in terms of peak and residual interstory drift ratios, and (2) the ground motion
duration plays an important role in the damage predicted when considering SSI effects, (3) in the
building models with SSI, a great amount of energy is dissipated by yielding of the zero-length
elements of the nonlinear SSI model, resulting on a significant reduction of the energy dissipated
by the beams of the buildings.

6.2 Main Contributions

 Manuscript 1:
 A new calibrated nonlinear FEM model was developed for the ELSA laboratory
tests performed on the SPEAR building example, which served as the basis for the
work performed in manuscript 1.
 A new ground motion selection procedure was proposed for isolating the effects
of duration in 3-D structural analysis, in which the spectral shape of bi-directional
components of short-duration ground motions are matched with bi-directional
components of long-duration ground motions, forming a set of spectrally
187

equivalent bi-directional set of scaled and rotated short-duration ground motions


for application in two orthogonal directions, for use in 3-D structures when
subjected to bi-directional ground motions.
 New collapse fragility curves based on DIP&A and DIR&V are proposed for this
type of structure to account for the effects of ground motion duration.
 Manuscript 2:
 New calibrated nonlinear FEM models were developed to represent:
o The lap-splice behavior of the bridge non-retrofitted columns.
o The solution adopted for retrofitting of the bridge column ends above the
footing using TiABs for longitudinal reinforcement and additional spirals.
 Manuscript 3:
 A new empirically calibrated phenomenological Q-y type multi-spring model is
implemented for use when modeling shallow foundations.

6.3 Overview of Main Findings

After analysis and discussion of the results obtained in the studies described above, the
following main conclusions are taken:

 The ground motion duration plays an important role in:


o The damage estimation of 3-D plan asymmetric RC buildings: increased ground
motion duration induces greater damage on the building.
o The seismic assessment of vintage as-built and retrofitted bridges: bridges
subjected to longer duration ground motions show greater levels of damage.
 The proposed retrofit solution applied to the columns of bridges provides a significant
increase to the structural strength and deformation capacity; however, the retrofit solution is
not as efficient when the bridge has columns with different lengths along its longitudinal
direction versus the idealized case when the bridge has columns with the same lengths;
 The inclusion of SSI effects increases measured response of peak and residual deformations
of the steel moment resisting frame building, while ground motion duration still plays an
important role in the damage predicted when considering SSI effects. A significant amount of
188

energy is absorbed by the soil, due to the yielding of the elements of the nonlinear SSI model
added the building models.

6.4 Future Work

As consequence of all the work performed in this dissertation, other questions arose and
remain unanswered. The following topics are therefore proposed to the research community:

 Assessment of ground motion duration effects of plan asymmetric ductile buildings


designed according to the current codes.
 Assessment of the ground motion duration effects on bridges considering a different
modeling approach for the abutments. The abutment cap beams should be modeled and
the impact of deck unseating and potential pile or abutment supports failures should be
considered.
 Assessment of ground motion duration effects through hazard-consistent ground motion
selection is worth focusing further on to study the effects of the conditional distribution
of ground motion duration when selecting site-specific ground motions.
 Assessment of ground motion duration effects on 3-D buildings/bridges taking into
account soil-structure interaction effects is worth further investigation. The
phenomenological models that incorporate the SSI effects can be extended to 3-D
analyses.
 The impact of the findings in this study on existing seismic risk methodologies that are
typically calibrated only to short-duration earthquake ground motions should be
investigated in the future.
189

APPENDICES

TABLE OF CONTENTS

Page

Tables with ground motions information ................................................................................... 196

Manuscript 3: Copulas with correlations between starting data and reduced data used for the
bearing pressure model .............................................................................................................. 205

Manuscript 3 - Code for implementation of HyperbolicQy material model ............................. 248

LIST OF FIGURES

Figure Page

Figure A 1 - Copulas showing the parameters correlations used for footing #2 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 205
Figure A 2 - Copulas showing the parameters correlations used for footing #3 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 206
Figure A 3 - Copulas showing the parameters correlations used for footing #4 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 207
Figure A 4 - Copulas showing the parameters correlations used for footing #5 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 208
Figure A 5 - Copulas showing the parameters correlations used for footing #1 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
190

m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 209
Figure A 6 - Copulas showing the parameters correlations used for footing #2 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 210
Figure A 7 - Copulas showing the parameters correlations used for footing #3 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 211
Figure A 8 - Copulas showing the parameters correlations used for footing #4 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 212
Figure A 9 - Copulas showing the parameters correlations used for footing #5 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 213
Figure A 10 - Copulas showing the parameters correlations used for footing #1 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 214
Figure A 11 - Copulas showing the parameters correlations used for footing #2 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 215
Figure A 12 - Copulas showing the parameters correlations used for footing #3 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 216
Figure A 13 - Copulas showing the parameters correlations used for footing #4 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
191

m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 217
Figure A 14 - Copulas showing the parameters correlations used for footing #5 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 218
Figure A 15 - Copulas showing the parameters correlations used for footing #1 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 219
Figure A 16 - Copulas showing the parameters correlations used for footing #2 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 220
Figure A 17 - Copulas showing the parameters correlations used for footing #3 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 221
Figure A 18 - Copulas showing the parameters correlations used for footing #4 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 222
Figure A 19 - Copulas showing the parameters correlations used for footing #5 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 223
Figure A 20 - Copulas showing the parameters correlations used for footing #1 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 224
Figure A 21 - Copulas showing the parameters correlations used for footing #2 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
192

m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 225
Figure A 22 - Copulas showing the parameters correlations used for footing #3 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 226
Figure A 23 - Copulas showing the parameters correlations used for footing #4 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 227
Figure A 24 - Copulas showing the parameters correlations used for footing #5 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 228
Figure A 25 - Copulas showing the parameters correlations used for footing #6 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 229
Figure A 26 - Copulas showing the parameters correlations used for footing #1 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 230
Figure A 27 - Copulas showing the parameters correlations used for footing #2 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 231
Figure A 28 - Copulas showing the parameters correlations used for footing #3 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 232
Figure A 29 - Copulas showing the parameters correlations used for footing #4 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
193

m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 233
Figure A 30 - Copulas showing the parameters correlations used for footing #5 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 234
Figure A 31 - Copulas showing the parameters correlations used for footing #6 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 235
Figure A 32 - Copulas showing the parameters correlations used for footing #1 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 236
Figure A 33 - Copulas showing the parameters correlations used for footing #2 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 237
Figure A 34 - Copulas showing the parameters correlations used for footing #3 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 238
Figure A 35 - Copulas showing the parameters correlations used for footing #4 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 239
Figure A 36 - Copulas showing the parameters correlations used for footing #5 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 240
Figure A 37 - Copulas showing the parameters correlations used for footing #6 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
194

m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 241
Figure A 38 - Copulas showing the parameters correlations used for footing #1 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 242
Figure A 39 - Copulas showing the parameters correlations used for footing #2 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 243
Figure A 40 - Copulas showing the parameters correlations used for footing #3 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 244
Figure A 41 - Copulas showing the parameters correlations used for footing #4 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 245
Figure A 42 - Copulas showing the parameters correlations used for footing #5 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 246
Figure A 43 - Copulas showing the parameters correlations used for footing #6 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC .......................................................................................................................... 247
195

LIST OF TABLES

Table Page

Table A 1 – Manuscript 1: Long-duration and short-duration ground motions used with


respective scale factor and rotation angle applied to the short-duration earthquake record. ...... 196
Table A 2 – Manuscript 1: Long-duration ground motion characteristics.................................. 197
Table A 3 – Manuscript 1: Short-duration ground motion characteristics ................................. 198
Table A 4 – Manuscript 2: Long-duration set and short-duration set (bridges RB1 and RB1R) of
the ground motions with respective scale factor and rotation angle applied to the short-duration
earthquake records ...................................................................................................................... 199
Table A 5 – Manuscript 2: Long-duration set and short-duration set (bridges RB2, RB2R, IB1,
IB1R, IB2, IB2R, IB3 and IB3R) of the ground motions with respective scale factor and rotation
angle applied ............................................................................................................................... 200
Table A 6 – Manuscript 2: Characteristics of the long-duration ground motions ...................... 201
Table A 7 – Manuscript 2: Characteristics of the short-duration ground motions selected for
bridges RB1 and RB1R ............................................................................................................... 202
Table A 8 – Manuscript 2: Characteristics of the short-duration ground motions selected to be
used in bridges RB2, RB2R, IB1, IB1R, IB2, IB2R, IB3 and IB3R .......................................... 203
Table A 9 – Manuscript 3: Long- and short duration ground motions selected and respective
scale factors applied in IDA analysis .......................................................................................... 204
Table A 1 – Manuscript 1: Long-duration and short-duration ground motions used with respective scale
factor and rotation angle applied to the short-duration earthquake record.

Pair Long-duration earthquake record Short-duration earthquake record θ SF


Number Earthquake Year Station Earthquake Year Station [degrees] [-] RMSEGM
1 Tohoku (Japan) 2011 Towadako Hector Mine (US) 1999 Hector 105 0.38 2.65
2 Tohoku (Japan) 2011 Nagawa El Mayor-Cucapah (Mexico) 2010 Michoacan de Ocampo 80 0.25 3.39
3 Tohoku (Japan) 2011 Takasato El Mayor-Cucapah (Mexico) 2010 Calexico Fire Station 165 0.47 2.48
4 Tohoku (Japan) 2011 Fukushima Chuetsu-oki (Japan) 2007 Hinodecho Yoshida Tsubame City 5 2.62 3.05
5 Tohoku (Japan) 2011 Ichinohe Northridge (US) 1994 LA - Brentwood VA Hospital 0 0.82 2.88
6 Tohoku (Japan) 2011 Shizukuishi Darfield (New Zeland) 2010 WSFC 0 2.18 3.38
7 Tohoku (Japan) 2011 Hanamaki Corinth (Greece) 1981 Corinth 5 1.75 3.37
8 Tohoku (Japan) 2011 Kanegasaki Irpinia (Italy) 1980 Rionero in Vulture 85 1.98 3.14
9 Tohoku (Japan) 2011 Ichinoseki Northridge (US) 1994 LA - Fletcher Dr 100 2.02 3.92
10 Tohoku (Japan) 2011 Iwanuma Chalfant Valley-02 (US) 1986 Long Valley Dam 175 3.33 3.67
11 Tohoku (Japan) 2011 Shiroishi Loma Prieta (US) 1989 Coyote Lake Dam 165 1.31 3.27
Tables with ground motions information

12 Tohoku (Japan) 2011 Shizugawa Northridge (US) 1994 LA - Century City CC North 165 0.86 4.64
13 Tohoku (Japan) 2011 Tendou Imperial Valley-06 (US) 1979 El Centro Array #12 0 1.41 3.21
14 Tohoku (Japan) 2011 Yamagata Big Bear-01 (US) 1992 Morongo Valley Fire Station 170 0.99 3.47
15 Maule (Chile) 2010 Constituition El Mayor-Cucapah (Mexico) 2010 El Centro - Imperial & Ross 100 1.69 3.61
16 Maule (Chile) 2010 Curico Chuetsu-oki (Japan) 2007 Sanjo Shinbori 100 1.31 3.58
17 Maule (Chile) 2010 Hualele Northridge (US) 1994 LA - W 15th St 100 3.48 3.34
18 Maule (Chile) 2010 Talca Imperial Valley-06 (US) 1979 Cerro Prieto 0 1.88 3.64
19 Valparaiso (Chile) 1985 Llolleo Chuetsu-oki (Japan) 2007 Joetsu Yasuzukaku Yasuzuka 165 2.40 2.34
20 Valparaiso (Chile) 1985 Valparaiso Elmandral Christchurch (New Zeland) 2011 Styx Mill Transfer Station 160 1.38 2.56
21 Valparaiso (Chile) 1985 San Isisdro Loma Prieta (US) 1989 Gilroy Array #7 90 2.83 3.41
22 Valparaiso (Chile) 1985 Vina del Mar Northridge (US) 1994 Hollywood - Willoughby Ave 100 1.34 3.71
23 ChiChi (Taiwan) 1999 CHY004 Northridge-01 (US) 1994 LA - Pico & Sentous 0 0.88 3.06
24 ChiChi (Taiwan) 1999 CHY076 Northridge-01 (US) 1994 Lawndale - Osage Ave 0 0.94 2.82
Imperial Valley Wildlife
25 ChiChi (Taiwan) 1999 CHY082 Superstition Hills-02 (US) 1987 20 0.32 2.74
Liquefaction Array
26 ChiChi (Taiwan) 1999 CHY107 Victoria (Mexico) 1980 Chihuahua 170 0.80 2.67
27 ChiChi AfSh (Taiwan) 1999 CHY012 Imperial Valley-06 (US) 1979 Westmorland Fire Station 0 0.75 2.35
28 Kocaeli (Turkey) 1999 Bursa Tofas Chuetsu-oki (Japan) 2007 Joetsu Kita 165 0.87 3.12
Fairbanks - Geophysic. Obs
29 Denali (Alaska) 2002 Chuetsu-oki (Japan) 2007 Joetsu Yanagishima Paddocks 0 1.03 3.72
CIGO,
30 Landers (US) 1992 Downey - Co Maint Bldg Parkfield-02 (US) 2004 Coaling - Priest Valley 85 1.93 2.78
31 Landers (US) 1992 Mission Creek Fault Northridge-01 (US) 1994 Inglewood - Union Oil 95 0.31 3.73
San Bernardino - 2nd &
32 Niigata (Japan) 2004 NIG010 Big Bear-01 (US) 1992 85 0.57
Arrowhead
196
197

Table A 2 – Manuscript 1: Long-duration ground motion characteristics


Rup.
Record D5-75 PGA PI Tp Sa(T1)
Earthquake Station Filename Mw Dist.
(#,component)
[sec] [km] [g] [sec] [g]
1H1 KR_AOMH121103111446EW2 53 0.11 0.00 1.6
Tohoku (Japan) Towadako KR_AOMH121103111446NS2
9.0 312 0.173
1H2 55 0.12 0.00 2.8
2H1 KR_AOMH171103111446EW2 53 0.14 0.00 1.3
Tohoku (Japan) Nagawa KR_AOMH171103111446NS2
9.0 292 0.204
2H2 53 0.11 0.00 3.4
3H1 KR_FKSH031103111446EW2 64 0.12 0.00 2.3
Tohoku (Japan) Takasato KR_FKSH031103111446NS2
9.0 279 0.273
3H2 66 0.13 0.00 2.8
4H1 KR_FKSH161103111446EW2 77 0.21 0.00 2.1
Tohoku (Japan) Fukushima KR_FKSH161103111446NS2
9.0 221 0.276
4H2 77 0.33 0.00 1.5
5H1 KR_IWTH111103111446EW2 52 0.14 0.00 1.2
Tohoku (Japan) Ichinohe KR_IWTH111103111446NS2
9.0 263 0.359
5H2 54 0.15 0.00 3.5
6H1 KR_IWTH161103111446EW2 61 0.12 0.01 1.5
Tohoku (Japan) Shizukuishi KR_IWTH161103111446NS2
9.0 238 0.325
6H2 61 0.16 0.00 1.6
7H1 KR_IWTH201103111446EW2 54 0.41 0.00 1.8
Tohoku (Japan) Hanamaki KR_IWTH201103111446NS2
9.0 209 0.610
7H2 54 0.37 0.00 1.3
8H1 KR_IWTH241103111446EW2 58 0.17 0.00 1.5
Tohoku (Japan) Kanegasaki KR_IWTH241103111446NS2
9.0 202 0.354
8H2 65 0.19 0.00 4.0
9H1 KR_IWTH261103111446EW2 56 0.53 0.00 1.3
Tohoku (Japan) Ichinoseki KR_IWTH261103111446NS2
9.0 188 0.707
9H2 54 0.52 0.00 1.1
10H1 KR_MYGH081103111446EW2 70 0.26 0.00 1.3
Tohoku (Japan) Iwanuma KR_MYGH081103111446NS2
9.0 177 0.415
10H2 66 0.29 0.50 2.2
11H1 KR_MYGH091103111446EW2 70 0.33 0.00 1.5
Tohoku (Japan) Shiroishi KR_MYGH091103111446NS2
9.0 198 0.638
11H2 70 0.32 0.00 2.1
12H1 KR_MYGH121103111446EW2 58 0.53 0.01 1.5
Tohoku (Japan) Shizugawa KR_MYGH121103111446NS2
9.0 137 0.271
12H2 56 0.46 0.02 2.4
13H1 KR_YMTH011103111446EW2 71 0.20 0.00 2.4
Tohoku (Japan) Tendou KR_YMTH011103111446NS2
9.0 219 0.214
13H2 64 0.18 0.00 2.4
14H1 KR_YMTH021103111446EW2 78 0.14 0.01 3.0
Tohoku (Japan) Yamagata KR_YMTH021103111446NS2 9.0 229 0.380
14H2 84 0.14 0.00 2.5
15H1 CONSTITUTION_EW 32 0.53 0.00 2.2
Maule (Chile) Constituition CONSTITUTION_NS
8.8 39 1.619
15H2 32 0.65 0.00 1.1
16H1 CURICO_EW 38 0.48 0.00 1.3
Maule (Chile) Curico CURICO_NS
8.8 65 0.587
16H2 37 0.41 0.00 1.9
17H1 HUALANE_EW 35 0.38 0.01 1.2
Maule (Chile) Hualele HUALANE_NS
8.8 50 0.856
17H2 34 0.48 0.00 1.4
18H1 TALCA_EW 51 0.46 0.00 0.9
Maule (Chile) Talca TALCA_NS
8.8 59 0.604
18H2 52 0.45 0.00 2.9
19H1 Valparaiso LLOLLEO_EW 28 0.39 0.00 1.6
Llolleo LLOLLEO_NS
7.8 N/A 0.924
19H2 (Chile) 27 0.68 0.00 1.1
20H1 Valparaiso Valparaiso VALPARAISO_EW 31 0.16 0.00 1.4
VALPARAISO_NS
7.8 N/A 0.629
20H2 (Chile) Elmandral 37 0.29 0.00 1.0
21H1 Valparaiso SANISIDRO_EW 30 0.70 0.00 0.4
San Isisdro SANISIDRO_NS
7.8 N/A 0.914
21H2 (Chile) 24 0.70 0.00 0.6
22H1 Valparaiso Vina del VINADELMAR_EW 32 0.32 0.00 1.0
VINADELMAR_NS
7.8 N/A 0.857
22H2 (Chile) Mar 32 0.22 0.02 1.1
23H1 ChiChi CHICHI_CHY004-W 36 0.10 0.00 6.4
CHY004 CHICHI_CHY004-N
7.62 47 0.169
23H2 (Taiwan) 30 0.10 0.00 7.6
24H1 ChiChi CHICHI_CHY076-E 31 0.07 0.57 6.3
CHY076 CHICHI_CHY076-n
7.62 42 0.162
24H2 (Taiwan) 26 0.07 0.00 7.7
25H1 ChiChi CHICHI_CHY082-E 26 0.06 0.02 7.9
CHY082 CHICHI_CHY082-N
7.62 36 0.122
25H2 (Taiwan) 30 0.06 0.01 7.8
26H1 ChiChi CHICHI_CHY107-W 32 0.09 0.09 4.6
CHY107 CHICHI_CHY107-N
7.62 51 0.230
26H2 (Taiwan) 27 0.10 0.00 5.6
27H1 ChiChi AfSh CHICHI.06_CHY012W 41 0.06 0.00 3.5
CHY012 CHICHI.06_CHY012N
6.3 89 0.105
27H2 (Taiwan) 32 0.04 0.00 2.7
28H1 Kocaeli KOCAELI_BUR000 26 0.10 0.01 8.0
Bursa Tofas KOCAELI_BUR090
7.51 60 0.293
28H2 (Turkey) 22 0.10 0.05 5.3
29H1 Geophysic. DENALI_FAIGO-90 22 0.09 0.00 3.4
Denali (Alaska) DENALI_FAIGO360 7.9 141 0.185
29H2 Obs CIGO, 28 0.07 0.00 7.2
30H1 Downey - LANDERS_DWN000 22 0.05 0.86 10.7
Landers (US) LANDERS_DWN090
7.28 158 0.095
30H2 Maint Bldg 29 0.04 0.02 6.1
31H1 Mission LANDERS_MCF000 23 0.13 0.00 4.0
Landers (US) LANDERS_MCF090
7.28 27 0.207
31H2 Creek Fault 31 0.13 0.01 5.2
32H1 NIIGATA_NIG010EW 11 0.11 0.01 4.7
Niigata (Japan) NIG010 NIIGATA_NIG010NS
6.63 58 0.108
32H2 29 0.07 0.00 6.7
198

Table A 3 – Manuscript 1: Short-duration ground motion characteristics


Record D5-75 Rup. Dist. PGA PI Tp Sa(T1)
Earthquake Station Filename Mw
(#,component) [sec] [km] [g] [sec] [g]
1H1 Hector Mine HECTOR_HEC000 8 0.27 0.47 1.48
Hector HECTOR_HEC090
7.13 10 0.174
1H2 (US) 7 0.33 0.43 7.03
2H1 El Mayor-Cuc. Michoacan de SIERRA.MEX_MDO000 22 0.54 0.03 1.74
SIERRA.MEX_MDO090
7.2 16 0.188
2H2 (Mexico) Ocampo 21 0.41 0.00 1.07
3H1 El Mayor-Cuc. Calexico Fire SIERRA.MEX_CXO090 18 0.27 0.07 5.36
SIERRA.MEX_CXO360
7.2 19 0.284
3H2 (Mexico) Station 19 0.26 0.00 7.04
4H1 Chuetsu-oki CHUETSU_65085EW 12 0.08 0.00 2.80
Hinodecho CHUETSU_65085NS
6.8 24 0.313
4H2 (Japan) 12 0.07 0.47 1.61
5H1 Northridge LABrentwood NORTHR_BVA195 6 0.19 0.32 2.13
NORTHR_BVA285
6.69 23 0.368
5H2 (US) VA Hospital 6 0.16 0.24 2.49
6H1 Darfield (New DARFIELD_WSFCN38W 15 0.07 0.05 5.82
WSFC DARFIELD_WSFCS52W
7 27 0.456
6H2 Zeland) 14 0.07 0.00 5.06
7H1 Corinth CORINTH_EW 6 0.24 0.02 2.35
Corinth CORINTH_NS
6.6 10 0.787
7H2 (Greece) 5 0.30 0.02 1.16
8H1 Rionero in ITALY_B-VLT000 11 0.10 0.01 1.67
Irpinia (Italy) ITALY_B-VLT270
6.2 23 0.371
8H2 Vulture 9 0.10 0.00 1.25
9H1 Northridge LA Fletcher NORTHRIDGE_FLE144 6 0.17 0.82 0.83
NORTHRIDGE_FLE234
6.69 27 1.028
9H2 (US) Dr 6 0.24 0.00 3.28
10H1 Chalfant Long Valley CHALFANT_LVL000 5 0.08 0.00 3.76
CHALFANT_LVL090 6.19 21 0.398
10H2 Valley-02 (US) Dam 8 0.07 0.22 1.20
11H1 Loma Prieta Coyote Lake LOMAP_CYC195 6 0.15 0.00 2.66
LOMAP_CYC285
6.93 20 1.088
11H2 (US) Dam 4 0.48 0.02 1.51
12H1 Northridge LA Century NORTHR_CCN090 7 0.26 0.33 1.43
NORTHR_CCN360
6.69 23 0.316
12H2 (US) City CC 7 0.22 0.00 2.35
13H1 Imperial El Centro IMPVALL_E12140 10 0.14 0.58 8.95
IMPVALL_E12230 6.53 18 0.277
13H2 Valley-06 (US) Array #12 10 0.12 0.01 6.38
14H1 Big Bear-01 Mor.Valley BIGBEAR_MVH045 8 0.15 0.58 1.74
BIGBEAR_MVH135
6.46 29 0.313
14H2 (US) Fire Sta. 8 0.12 0.01 1.07
15H1 El Mayor-Cuc. El Centro SIERRA.MEX01711_90 15 0.37 0.00 4.47
SIERRA.MEX01711_360
7.2 19 1.120
15H2 (Mexico) Imp. & Ross 15 0.38 0.00 8.31
16H1 Chuetsu-oki Sanjo CHUETSU_65033EW 10 0.26 0.00 1.48
CHUETSU_65033NS
6.8 23 0.752
16H2 (Japan) Shinbori 11 0.32 0.00 4.68
17H1 Northridge LA - W 15th NORTHR_W15090 9 0.10 0.00 2.67
NORTHR_W15180
6.69 30 0.720
17H2 (US) St 10 0.17 0.00 4.17
18H1 Imperial IMPVALL_CPE147 17 0.17 0.00 5.31
Cerro Prieto IMPVALL_CPE237
6.53 15 0.806
18H2 Valley-06 (US) 20 0.16 0.00 4.84
19H1 Chuetsu-oki Joetsu CHUETSU_65004EW 6 0.22 0.11 1.05
CHUETSU_65004NS
6.8 42 0.480
19H2 (Japan) Yasuzukaku 5 0.15 0.03 0.90
20H1 Christchurch Styx Mill CCHURCH_S02W 5 0.18 0.08 1.36
CCHURCH_N88W
6.2 11 0.450
20H2 (New Zealand) Transfer Sta. 7 0.14 0.00 3.70
21H1 Loma Prieta Gilroy Array LOMAP_GMR000 4 0.26 0.00 0.61
LOMAP_GMR090
6.93 23 1.162
21H2 (US) #7 5 0.22 0.00 5.11
22H1 Northridge Willoughby NORTHR_WIL090 6 0.14 0.75 1.22
NORTHR_WIL180
6.69 23 0.605
22H2 (US) Ave 6 0.25 0.01 2.34
23H1 Northridge-01 LA - Pico & NORTHR_PIC090 9 0.19 0.16 0.85
NORTHR_PIC180
6.69 31 0.206
23H2 (US) Sentous 7 0.10 0.00 4.24
24H1 Northridge-01 Lawndale - NORTHR_LOA092 14 0.08 0.00 2.11
NORTHR_LOA182
6.69 40 0.140
24H2 (US) Osage Ave 14 0.15 0.00 2.88
25H1 Superstition Imp. Valley SUPER.B_B-IVW090 20 0.18 0.01 6.40
SUPER.B_B-IVW360
6.54 24 0.133
25H2 Hills-02 (US) Wildlie Array 15 0.21 0.00 6.35
26H1 Victoria VICT_CHI102 8 0.10 0.00 4.05
Chihuahua VICT_CHI192
6.33 19 0.308
26H2 (Mexico) 10 0.15 0.00 5.01
27H1 Imperial Westmorland IMPV.H_H-WSM090 13 0.11 0.44 3.54
IMPVL.H_H-WSM180
6.53 15 0.114
27H2 Valley-06 (US) Fire Station 14 0.08 0.00 6.38
28H1 Chuetsu-oki CHUETSU_65009EW 11 0.09 0.00 2.62
Joetsu Kita CHUETSU_65009NS
6.8 30 0.252
28H2 (Japan) 19 0.18 0.01 1.54
29H1 Chuetsu-oki Yanagishima CHUETSU_65003EW 3 0.33 0.00 7.28
CHUETSU_65003NS
6.8 31 0.144
29H2 (Japan) paddocks 4 0.28 0.00 0.63
30H1 Parkfield-02 Coaling - PARK2004_46174-90 10 0.02 0.01 7.55
PARK2004_46174360
6 22 0.107
30H2 (US) Priest Valley 9 0.03 0.00 2.20
31H1 Northridge-01 Inglewood - NORTHR_ING000 12 0.10 0.00 3.08
NORTHR_ING090
6.69 42 0.176
31H2 (US) Union Oil 10 0.09 0.02 2.58
32H1 Big Bear-01 San BIGBEAR_SB2270 10 0.11 0.00 2.17
BIGBEAR_SB2360
6.46 34 0.100
32H2 (US) Bernardino 14 0.10 0.00 2.73
Table A 4 – Manuscript 2: Long-duration set and short-duration set (bridges RB1 and RB1R) of the ground
motions with respective scale factor and rotation angle applied to the short-duration earthquake records

Pair Long-duration earthquake record Short-duration earthquake record θ SF


Number Earthquake Year Station Earthquake Year Station [degrees] [-]
1 Tohoku (Japan) 2011 Towadako Parkfield-02_ CA 2004 COALINGA - PRIEST VALLEY 100 0.38
El Mayor-Cucapah_
2 Tohoku (Japan) 2011 Nagawa 2010 El Centro - Imperial & Ross 80 0.26
Mexico
3 Tohoku (Japan) 2011 Takasato Northridge-01 1994 LA - Century City CC North 105 1.07
4 Tohoku (Japan) 2011 Fukushima Hector Mine 1999 Hector 90 2.71
Coyote Lake Dam - Southwest
5 Tohoku (Japan) 2011 Ichinohe Loma Prieta 1989 0 0.87
Abutment
6 Tohoku (Japan) 2011 Shizukuishi Coalinga-01 1983 Parkfield - Fault Zone 15 105 0.72
7 Tohoku (Japan) 2011 Hanamaki Big Bear-01 1992 Morongo Valley Fire Station 10 1.65
8 Tohoku (Japan) 2011 Kanegasaki Victoria_ Mexico 1980 Chihuahua 175 3.16
9 Tohoku (Japan) 2011 Ichinoseki Corinth_ Greece 1981 Corinth 90 1.6
10 Tohoku (Japan) 2011 Iwanuma Northern Calif-03 1954 Ferndale City Hall 105 0.73
11 Tohoku (Japan) 2011 Shiroishi Northridge-01 1994 Hollywood - Willoughby Ave 5 1.38
12 Tohoku (Japan) 2011 Shizugawa Chuetsu-oki_ Japan 2007 Sanjo Shinbori 160 0.82
13 Tohoku (Japan) 2011 Tendou Chuetsu-oki_ Japan 2007 Joetsu Kita 0 1.42
El Mayor-Cucapah_
14 Tohoku (Japan) 2011 Yamagata 2010 MICHOACAN DE OCAMPO 5 0.38
Mexico
15 Maule (Chile) 2010 Constituition Chuetsu-oki_ Japan 2007 Joetsu Yasuzukaku Yasuzuka 170 2.4
16 Maule (Chile) 2010 Curico Chalfant Valley-02 1986 Long Valley Dam (L Abut) 85 3.77
17 Maule (Chile) 2010 Hualele Morgan Hill 1984 Hollister Differential Array #3 80 1.57
18 Maule (Chile) 2010 Talca Loma Prieta 1989 Anderson Dam (Downstream) 175 1.7
19 Valparaiso (Chile) 1985 Llolleo Duzce_ Turkey 1999 Bolu 0 2.27
20 Valparaiso (Chile) 1985 Valparaiso Elmandral Joshua Tree_ CA 1992 Indio - Jackson Road 165 1.37
21 Valparaiso (Chile) 1985 San Isisdro Loma Prieta 1989 Gilroy Array #7 175 2.31
22 Valparaiso (Chile) 1985 Vina del Mar Joshua Tree_ CA 1992 Morongo Valley Fire Station 85 2.12
Imperial Valley Wildlife Liquefaction
23 ChiChi (Taiwan) 1999 CHY004 Superstition Hills-02 1987 95 0.39
Array
24 ChiChi (Taiwan) 1999 CHY076 Darfield_ New Zealand 2010 Styx Mill Transfer Station 0 2.06
25 ChiChi (Taiwan) 1999 CHY082 Parkfield-02_ CA 2004 Parkfield - Fault Zone 3 80 0.55
26 ChiChi (Taiwan) 1999 CHY107 San Simeon_ CA 2003 San Luis Obispo - Lopez Lake Grounds 80 0.82
ChiChi AfSh
27 1999 CHY012 Imperial Valley-06 1979 Westmorland Fire Sta 95 0.3
(Taiwan)
28 Kocaeli (Turkey) 1999 Bursa Tofas Tabas_ Iran 1978 Boshrooyeh 95 1.02
Fairbanks -
29 Denali (Alaska) 2002 Geophysic. Obs Chuetsu-oki_ Japan 2007 Joetsu Yanagishima paddocks 10 0.26
CIGO,
Downey - Co Maint
30 Landers (US) 1992 Iwate_ Japan 2008 Yuzawa Town 85 0.28
Bldg
31 Landers (US) 1992 Mission Creek Fault Darfield_ New Zealand 2010 Christchurch Cashmere High School 95 0.76
32 Niigata (Japan) 2004 NIG010 Coalinga-01 1983 Parkfield - Stone Corral 4E 85 0.63
199
Pair Long-duration earthquake record Short-duration earthquake record θ SF
Number Earthquake Year Earthquake Year Earthquake Year [degrees] [-]
1 Tohoku (Japan) 2011 Towadako Imperial Valley-06 1979 El Centro Array #12 100 0.38
2 Tohoku (Japan) 2011 Nagawa El Mayor-Cucapah_ Mexico 2010 El Centro - Imperial & Ross 80 0.26
3 Tohoku (Japan) 2011 Takasato El Mayor-Cucapah_ Mexico 2010 Calexico Fire Station 105 1.07
4 Tohoku (Japan) 2011 Fukushima Darfield_ New Zealand 2010 Christchurch Cashmere High School 90 2.71
5 Tohoku (Japan) 2011 Ichinohe Northridge-01 1994 LA - Brentwood VA Hospital 0 0.87
6 Tohoku (Japan) 2011 Shizukuishi Darfield_ New Zealand 2010 WSFC 105 0.72
7 Tohoku (Japan) 2011 Hanamaki Northridge-01 1994 Canyon Country - W Lost Cany 10 1.65
8 Tohoku (Japan) 2011 Kanegasaki Northridge-01 1994 LA - W 15th St 175 3.16
9 Tohoku (Japan) 2011 Ichinoseki Northridge-01 1994 LA - N Faring Rd 90 1.60
10 Tohoku (Japan) 2011 Iwanuma Superstition Hills-02 1987 El Centro Imp. Co. Cent 105 0.73
11 Tohoku (Japan) 2011 Shiroishi Northridge-01 1994 Sunland - Mt Gleason Ave 5 1.38
12 Tohoku (Japan) 2011 Shizugawa Northridge-01 1994 LA - Century City CC North 160 0.82
13 Tohoku (Japan) 2011 Tendou Loma Prieta 1989 Sunnyvale - Colton Ave. 0 1.42
14 Tohoku (Japan) 2011 Yamagata El Mayor-Cucapah_ Mexico 2010 MICHOACAN DE OCAMPO 5 0.38
15 Maule (Chile) 2010 Constituition Chuetsu-oki_ Japan 2007 Joetsu Yasuzukaku Yasuzuka 170 2.40
16 Maule (Chile) 2010 Curico Umbria Marche_ Italy 1997 Castelnuovo-Assisi 85 3.77
17 Maule (Chile) 2010 Hualele Morgan Hill 1984 Hollister Differential Array #3 80 1.57
18 Maule (Chile) 2010 Talca Imperial Valley-06 1979 Cerro Prieto 175 1.70
19 Valparaiso (Chile) 1985 Llolleo Basso Tirreno_ Italy 1978 Patti-Cabina Prima 0 2.27
Valparaiso
20 Valparaiso (Chile) 1985 N. Palm Springs 1986 Whitewater Trout Farm 165 1.37
Elmandral
21 Valparaiso (Chile) 1985 San Isisdro Joshua Tree_ CA 1992 Indio - Jackson Road 175 2.31
angle applied

22 Valparaiso (Chile) 1985 Vina del Mar Spitak_ Armenia 1988 Gukasian 85 2.12
23 ChiChi (Taiwan) 1999 CHY004 Victoria_ Mexico 1980 Chihuahua 95 0.39
24 ChiChi (Taiwan) 1999 CHY076 Darfield_ New Zealand 2010 Styx Mill Transfer Station 0 2.06
25 ChiChi (Taiwan) 1999 CHY082 Iwate_ Japan 2008 Yuzawa Town 80 0.55
26 ChiChi (Taiwan) 1999 CHY107 Chuetsu-oki_ Japan 2007 Joetsu Kita 80 0.82
ChiChi AfSh
27 1999 CHY012 Imperial Valley-06 1979 Westmorland Fire Sta 95 0.30
(Taiwan)
28 Kocaeli (Turkey) 1999 Bursa Tofas Chuetsu-oki_ Japan 2007 Sanjo Shinbori 95 1.02
Fairbanks -
29 Denali (Alaska) 2002 Geophysic. Obs Chuetsu-oki_ Japan 2007 Joetsu Yanagishima paddocks 10 0.26
CIGO,
Downey - Co
30 Landers (US) 1992 Parkfield-02_ CA 2004 COALINGA - PRIEST VALLEY 85 0.28
Maint Bldg
Mission Creek
31 Landers (US) 1992 Northridge-01 1994 Inglewood - Union Oil 95 0.76
Fault
32 Niigata (Japan) 2004 NIG010 Big Bear-01 1992 San Bernardino - 2nd & Arrowhead 85 0.63
Table A 5 – Manuscript 2: Long-duration set and short-duration set (bridges RB2, RB2R, IB1,
IB1R, IB2, IB2R, IB3 and IB3R) of the ground motions with respective scale factor and rotation
200
201

Table A 6 – Manuscript 2: Characteristics of the long-duration ground motions


Record D5-75 Rup. Dist PGA Sa(T1)
Earthquake Station Filename Mw
(#,component) [sec] [km]. [g] [g]
1H1 KR_AOMH121103111446EW2 53 0.11
1H2
Tohoku_Japan Towadako
KR_AOMH121103111446NS2 55
9 312
0.12
[g]
2H1 KR_AOMH171103111446EW2 53 0.14
Tohoku_Japan Nagawa 9 292 0.204
2H2 KR_AOMH171103111446NS2 53 0.11
3H1 KR_FKSH031103111446EW2 64 0.12
Tohoku_Japan Takasato 9 279 0.273
3H2 KR_FKSH031103111446NS2 66 0.13
4H1 KR_FKSH161103111446EW2 77 0.21
Tohoku_Japan Fukushima 9 221 0.276
4H2 KR_FKSH161103111446NS2 77 0.33
5H1 KR_IWTH111103111446EW2 52 0.14
Tohoku_Japan Ichinohe 9 263 0.359
5H2 KR_IWTH111103111446NS2 54 0.15
6H1 KR_IWTH161103111446EW2 61 0.12
Tohoku_Japan Shizukuishi 9 238 0.325
6H2 KR_IWTH161103111446NS2 61 0.16
7H1 KR_IWTH201103111446EW2 54 0.41
Tohoku_Japan Hanamaki 9 209 0.61
7H2 KR_IWTH201103111446NS2 54 0.37
8H1 KR_IWTH241103111446EW2 58 0.17
Tohoku_Japan Kanegasaki 9 202 0.354
8H2 KR_IWTH241103111446NS2 65 0.19
9H1 KR_IWTH261103111446EW2 56 0.53
Tohoku_Japan Ichinoseki 9 188 0.707
9H2 KR_IWTH261103111446NS2 54 0.52
10H1 KR_MYGH081103111446EW2 70 0.26
Tohoku_Japan Iwanuma 9 177 0.415
10H2 KR_MYGH081103111446NS2 66 0.29
11H1 KR_MYGH091103111446EW2 70 0.33
Tohoku_Japan Shiroishi 9 198 0.638
11H2 KR_MYGH091103111446NS2 70 0.32
12H1 KR_MYGH121103111446EW2 58 0.53
Tohoku_Japan Shizugawa 9 137 0.271
12H2 KR_MYGH121103111446NS2 56 0.46
13H1 KR_YMTH011103111446EW2 71 0.2
Tohoku_Japan Tendou 9 219 0.214
13H2 KR_YMTH011103111446NS2 64 0.18
14H1 KR_YMTH021103111446EW2 78 0.14
Tohoku_Japan Yamagata 9 229 0.38
14H2 KR_YMTH021103111446NS2 84 0.14
15H1 CONSTITUTION_EW 32 0.53
Maule_Chile Constituition 8.8 39 1.619
15H2 CONSTITUTION_NS 32 0.65
16H1 CURICO_EW 38 0.48
Maule_Chile Curico 8.8 65 0.587
16H2 CURICO_NS 37 0.41
17H1 HUALANE_EW 35 0.38
Maule_Chile Hualele 8.8 50 0.856
17H2 HUALANE_NS 34 0.48
18H1 TALCA_EW 51 0.46
Maule_Chile Talca 8.8 59 0.604
18H2 TALCA_NS 52 0.45
19H1 LLOLLEO_EW 28 0.39
Valparaiso_Chile Llolleo 7.8 N/A 0.924
19H2 LLOLLEO_NS 27 0.68
20H1 Valparaiso VALPARAISO_EW 31 0.16
Valparaiso_Chile 7.8 N/A 0.629
20H2 Elmandral VALPARAISO_NS 37 0.29
21H1 SANISIDRO_EW 30 0.7
Valparaiso_Chile San Isisdro 7.8 N/A 0.914
21H2 SANISIDRO_NS 24 0.7
22H1 Vina del VINADELMAR_EW 32 0.32
Valparaiso_Chile 7.8 N/A 0.857
22H2 Mar VINADELMAR_NS 32 0.22
23H1 CHICHI_CHY004-W 36 0.1
ChiChi_Taiwan CHY004 7.62 47 0.169
23H2 CHICHI_CHY004-N 30 0.1
24H1 CHICHI_CHY076-E 31 0.07
ChiChi_Taiwan CHY076 7.62 42 0.162
24H2 CHICHI_CHY076-n 26 0.07
25H1 CHICHI_CHY082-E 26 0.06
ChiChi_Taiwan CHY082 7.62 36 0.122
25H2 CHICHI_CHY082-N 30 0.06
26H1 CHICHI_CHY107-W 32 0.09
ChiChi_Taiwan CHY107 7.62 51 0.23
26H2 CHICHI_CHY107-N 27 0.1
27H1 ChiChi CHICHI.06_CHY012W 41 0.06
CHY012 6.3 89 0.105
27H2 AfSh_Taiwan CHICHI.06_CHY012N 32 0.04
28H1 KOCAELI_BUR000 26 0.1
Kocaeli_Turkey Bursa Tofas 7.51 60 0.293
28H2 KOCAELI_BUR090 22 0.1
29H1 Geophysic. DENALI_FAIGO-90 22 0.09
Denali_Alaska 7.9 141 0.185
29H2 Obs CIGO, DENALI_FAIGO360 28 0.07
30H1 Downey - LANDERS_DWN000 22 0.05
Landers 7.28 158 0.095
30H2 Maint Bldg LANDERS_DWN090 29 0.04
31H1 Mission LANDERS_MCF000 23 0.13
Landers 7.28 27 0.207
31H2 Creek Fault LANDERS_MCF090 31 0.13
32H1 NIIGATA_NIG010EW 11 0.11
Niigata_Japan NIG010 6.63 58 0.108
32H2 NIIGATA_NIG010NS 29 0.07
202

Table A 7 – Manuscript 2: Characteristics of the short-duration ground motions selected for


bridges RB1 and RB1R
Record D5-75 Rup. Dist PGA Sa(T1)
Earthquake Station Filename Mw
(#,component) [sec] [km]. [g] [g]
1H1 COALINGA - RSN4150_PARK2004_46174-90.AT2 9 0.02
Parkfield-02_ CA PRIEST 6 22.02 0.05
1H2 RSN4150_PARK2004_46174360.AT2 10 0.03
VALLEY
2H1 El Mayor- El Centro - RSN5837_SIERRA.MEX_01711360.AT2 15 0.37
7.2 20.08 0.41
2H2 Cucapah_ Mexico Imperial & Ross RSN5837_SIERRA.MEX_01711-90.AT2 15 0.38
3H1 LA - Century City RSN988_NORTHR_CCN090.AT2 7 0.26
Northridge-01 6.69 23.41 0.25
3H2 CC North RSN988_NORTHR_CCN360.AT2 7 0.22
4H1 RSN1787_HECTOR_HEC000.AT2 6 0.27
Hector Mine Hector 7.13 11.66 0.45
4H2 RSN1787_HECTOR_HEC090.AT2 8 0.33
5H1 Coyote Lake Dam RSN755_LOMAP_CYC195.AT2 6 0.15
Loma Prieta - Southwest 6.93 20.34 0.20
5H2 RSN755_LOMAP_CYC285.AT2 4 0.48
Abutment
6H1 Parkfield - Fault RSN339_COALINGA.H_H-Z15000.AT2 5 0.19
Coalinga-01 6.36 29.38 0.17
6H2 Zone 15 RSN339_COALINGA.H_H-Z15090.AT2 10 0.12
7H1 Morongo Valley RSN6059_BIGBEAR_MVH045.AT2 8 0.15
Big Bear-01 6.46 29.06 0.19
7H2 Fire Station RSN6059_BIGBEAR_MVH135.AT2 8 0.12
8H1 RSN266_VICT_CHI102.AT2 8 0.15
Victoria_ Mexico Chihuahua 6.33 18.96 0.20
8H2 RSN266_VICT_CHI192.AT2 11 0.10
9H1 RSN313_CORINTH_COR--L.AT2 5 0.24
Corinth_ Greece Corinth 6.6 10.27 0.18
9H2 RSN313_CORINTH_COR--T.AT2 5 0.30
10H1 RSN20_NCALIF.FH_H-FRN044.AT2 7 0.16
Northern Calif-03 Ferndale City Hall 6.5 27.02 0.35
10H2 RSN20_NCALIF.FH_H-FRN314.AT2 6 0.20
11H1 Hollywood - RSN978_NORTHR_WIL090.AT2 7 0.14
Northridge-01 6.69 23.07 0.22
11H2 Willoughby Ave RSN978_NORTHR_WIL180.AT2 6 0.25
12H1 Chuetsu-oki_ RSN4860_CHUETSU_65033NS.AT2 11 0.26
Sanjo Shinbori 6.8 23.18 0.35
12H2 Japan RSN4860_CHUETSU_65033EW.AT2 10 0.32
13H1 Chuetsu-oki_ RSN4840_CHUETSU_65003NS.AT2 11 0.18
Joetsu Kita 6.8 29.45 0.21
13H2 Japan RSN4840_CHUETSU_65003EW.AT2 20 0.09
14H1 El Mayor- MICHOACAN RSN5827_SIERRA.MEX_MDO000.AT2 20 0.54
7.2 15.91 0.46
14H2 Cucapah_ Mexico DE OCAMPO RSN5827_SIERRA.MEX_MDO090.AT2 23 0.41
15H1 Joetsu RSN4841_CHUETSU_65004NS.AT2 6 0.22
Chuetsu-oki_
Yasuzukaku 6.8 25.52 0.15
15H2 Japan RSN4841_CHUETSU_65004EW.AT2 7 0.15
Yasuzuka
16H1 Chalfant Valley- Long Valley Dam RSN554_CHALFANT.A_A-LVL000.AT2 5 0.08
6.19 21.12 0.10
16H2 02 (L Abut) RSN554_CHALFANT.A_A-LVL090.AT2 8 0.07
17H1 Hollister RSN464_MORGAN_HD3255.AT2 12 0.08
Morgan Hill Differential Array 6.19 26.43 0.13
17H2 RSN464_MORGAN_HD3345.AT2 10 0.08
#3
18H1 Anderson Dam RSN739_LOMAP_AND250.AT2 5 0.25
Loma Prieta 6.93 20.26 0.16
18H2 (Downstream) RSN739_LOMAP_AND340.AT2 5 0.24
19H1 RSN1602_DUZCE_BOL000.AT2 3 0.74
Duzce_ Turkey Bolu 7.14 12.04 0.53
19H2 RSN1602_DUZCE_BOL090.AT2 1 0.80
20H1 Indio - Jackson RSN6877_JOSHUA_5294180.AT2 3 0.41
Joshua Tree_ CA 6.1 25.53 0.27
20H2 Road RSN6877_JOSHUA_5294090.AT2 5 0.21
21H1 RSN993_NORTHR_FLE144.AT2 5 0.23
Loma Prieta Gilroy Array #7 6.93 22.68 0.07
21H2 RSN993_NORTHR_FLE234.AT2 4 0.32
22H1 Morongo Valley RSN6875_JOSHUA_5071045.AT2 6 0.13
Joshua Tree_ CA 6.1 22.3 0.13
22H2 Fire Station RSN6875_JOSHUA_5071135.AT2 8 0.07
23H1 Imperial Valley RSN729_SUPER.B_B-IVW090.AT2 19 0.18
Superstition Wildlife
6.54 23.85 0.31
23H2 Hills-02 Liquefaction RSN729_SUPER.B_B-IVW360.AT2 13 0.21
Array
24H1 Darfield_ New Styx Mill Transfer RSN6969_DARFIELD_SMTCN88W.AT2 11 0.18
7 20.86 0.19
24H2 Zealand Station RSN6969_DARFIELD_SMTCS02W.AT2 14 0.17
25H1 Parkfield - Fault RSN4108_PARK2004_COH090.AT2 3 0.38
Parkfield-02_ CA 6 2.73 0.18
25H2 Zone 3 RSN4108_PARK2004_COH360.AT2 4 0.40
26H1 San Luis Obispo - RSN3994_SANSIMEO_36153090.AT2 8 0.13
San Simeon_ CA Lopez Lake 6.52 48.11 0.08
26H2 RSN3994_SANSIMEO_36153360.AT2 10 0.12
Grounds
27H1 Imperial Valley- Westmorland Fire RSN192_IMPVALL.H_H-WSM090.AT2 14 0.08
6.53 15.25 0.07
27H2 06 Sta RSN192_IMPVALL.H_H-WSM180.AT2 13 0.11
28H1 RSN138_TABAS_BOS-L1.AT2 15 0.11
Tabas_ Iran Boshrooyeh 7.35 28.79 0.14
28H2 RSN138_TABAS_BOS-T1.AT2 15 0.08
29H1 Joetsu RSN4846_CHUETSU_65009NS.AT2 4 0.28
Chuetsu-oki_
Yanagishima 6.8 31.43 0.13
29H2 Japan RSN4846_CHUETSU_65009EW.AT2 3 0.33
paddocks
30H1 RSN5806_IWATE_55461NS.AT2 8 0.19
Iwate_ Japan Yuzawa Town 6.9 25.56 0.31
30H2 RSN5806_IWATE_55461EW.AT2 9 0.24
31H1 Christchurch RSN6890_DARFIELD_CMHSN10E.AT2 8 0.23
Darfield_ New
Cashmere High 7 17.64 0.23
31H2 Zealand RSN6890_DARFIELD_CMHSS80E.AT2 9 0.25
School
32H1 Parkfield - Stone RSN358_COALINGA.H_H-SC4000.AT2 5 0.07
Coalinga-01 6.36 31.58 0.12
32H2 Corral 4E RSN358_COALINGA.H_H-SC4090.AT2 6 0.07
203

Table A 8 – Manuscript 2: Characteristics of the short-duration ground motions selected to be


used in bridges RB2, RB2R, IB1, IB1R, IB2, IB2R, IB3 and IB3R
Record D5-75 Rup. Dist PGA Sa(T1)
Earthquake Station Filename Mw
(#,component) [sec] [km]. [g] [g]
1H1 Imperial Valley- El Centro Array RSN175_IMPVALL.H_H-E12140.AT2 10 0.14
6.53 17.94 0.16
1H2 06 #12 RSN175_IMPVALL.H_H-E12230.AT2 10 0.12
2H1 El Mayor- El Centro - RSN5837_SIERRA.MEX_01711360.AT2 15 0.37
7.2 20.08 0.87
2H2 Cucapah_ Mexico Imperial & Ross RSN5837_SIERRA.MEX_01711-90.AT2 15 0.38
3H1 El Mayor- Calexico Fire RSN5975_SIERRA.MEX_CXO360.AT2 18 0.27
7.2 20.46 0.43
3H2 Cucapah_ Mexico Station RSN5975_SIERRA.MEX_CXO090.AT2 19 0.26
4H1 Christchurch RSN6890_DARFIELD_CMHSN10E.AT2 8 0.23
Darfield_ New
Cashmere High 7 17.64 0.30
4H2 Zealand RSN6890_DARFIELD_CMHSS80E.AT2 9 0.25
School
5H1 LA - Brentwood RSN986_NORTHR_BVA195.AT2 6 0.19
Northridge-01 6.69 22.5 0.33
5H2 VA Hospital RSN986_NORTHR_BVA285.AT2 6 0.16
6H1 Darfield_ New RSN6988_DARFIELD_WSFCN38W.AT2 15 0.07
WSFC 7 26.93 0.11
6H2 Zealand RSN6988_DARFIELD_WSFCS52W.AT2 14 0.07
7H1 Canyon Country RSN960_NORTHR_LOS000.AT2 3 0.40
Northridge-01 6.69 12.44 0.52
7H2 - W Lost Cany RSN960_NORTHR_LOS270.AT2 3 0.47
8H1 RSN1008_NORTHR_W15090.AT2 10 0.10
Northridge-01 LA - W 15th St 6.69 29.74 0.21
8H2 RSN1008_NORTHR_W15180.AT2 9 0.17
9H1 RSN996_NORTHR_FAR000.AT2 5 0.28
Northridge-01 LA - N Faring Rd 6.69 20.81 0.33
9H2 RSN996_NORTHR_FAR090.AT2 6 0.26
10H1 Superstition Hills- El Centro Imp. RSN721_SUPER.B_B-ICC000.AT2 7 0.36
6.54 18.2 0.29
10H2 02 Co. Cent RSN721_SUPER.B_B-ICC090.AT2 9 0.26
11H1 Sunland - Mt RSN1083_NORTHR_GLE170.AT2 7 0.13
Northridge-01 6.69 13.35 0.26
11H2 Gleason Ave RSN1083_NORTHR_GLE260.AT2 5 0.16
12H1 LA - Century RSN988_NORTHR_CCN090.AT2 7 0.26
Northridge-01 6.69 23.41 0.36
12H2 City CC North RSN988_NORTHR_CCN360.AT2 7 0.22
13H1 Sunnyvale - RSN806_LOMAP_SVL270.AT2 10 0.21
Loma Prieta 6.93 24.23 0.27
13H2 Colton Ave. RSN806_LOMAP_SVL360.AT2 10 0.21
14H1 El Mayor- MICHOACAN RSN5827_SIERRA.MEX_MDO000.AT2 20 0.54
7.2 15.91 0.80
14H2 Cucapah_ Mexico DE OCAMPO RSN5827_SIERRA.MEX_MDO090.AT2 23 0.41
15H1 Joetsu RSN4841_CHUETSU_65004NS.AT2 6 0.22
Chuetsu-oki_
Yasuzukaku 6.8 25.52 0.35
15H2 Japan RSN4841_CHUETSU_65004EW.AT2 7 0.15
Yasuzuka
16H1 Umbria Marche_ Castelnuovo- RSN4348_UBMARCHE.P_A-CSA000.AT2 8 0.17
6 17.28 0.15
16H2 Italy Assisi RSN4348_UBMARCHE.P_A-CSA270.AT2 9 0.11
17H1 Hollister RSN464_MORGAN_HD3255.AT2 12 0.08
Morgan Hill Differential 6.19 26.43 0.15
17H2 RSN464_MORGAN_HD3345.AT2 10 0.08
Array #3
18H1 Imperial Valley- RSN164_IMPVALL.H_H-CPE147.AT2 17 0.17
Cerro Prieto 6.53 15.19 0.26
18H2 06 RSN164_IMPVALL.H_H-CPE237.AT2 20 0.16
19H1 Basso Tirreno_ Patti-Cabina RSN4285_BTIRRENO.P_PTT000.AT2 5 0.07
6 17.4 0.13
19H2 Italy Prima RSN4285_BTIRRENO.P_PTT090.AT2 4 0.16
20H1 Whitewater Trout RSN540_PALMSPR_WWT180.AT2 2 0.48
N. Palm Springs 6.06 6.04 0.40
20H2 Farm RSN540_PALMSPR_WWT270.AT2 2 0.63
21H1 Indio - Jackson RSN6877_JOSHUA_5294180.AT2 3 0.41
Joshua Tree_ CA 6.1 25.53 0.33
21H2 Road RSN6877_JOSHUA_5294090.AT2 5 0.21
22H1 RSN730_SPITAK_GUK000.AT2 6 0.20
Spitak_ Armenia Gukasian 6.77 23.99 0.29
22H2 RSN730_SPITAK_GUK090.AT2 4 0.17
23H1 RSN266_VICT_CHI102.AT2 8 0.15
Victoria_ Mexico Chihuahua 6.33 18.96 0.22
23H2 RSN266_VICT_CHI192.AT2 11 0.10
24H1 Darfield_ New Styx Mill RSN6969_DARFIELD_SMTCN88W.AT2 11 0.18
7 20.86 0.31
24H2 Zealand Transfer Station RSN6969_DARFIELD_SMTCS02W.AT2 14 0.17
25H1 RSN5806_IWATE_55461NS.AT2 8 0.19
Iwate_ Japan Yuzawa Town 6.9 25.56 0.25
25H2 RSN5806_IWATE_55461EW.AT2 9 0.24
26H1 Chuetsu-oki_ RSN4840_CHUETSU_65003NS.AT2 11 0.18
Joetsu Kita 6.8 29.45 0.21
26H2 Japan RSN4840_CHUETSU_65003EW.AT2 20 0.09
27H1 Imperial Valley- Westmorland RSN192_IMPVALL.H_H-WSM090.AT2 14 0.08
6.53 15.25 0.11
27H2 06 Fire Sta RSN192_IMPVALL.H_H-WSM180.AT2 13 0.11
28H1 Chuetsu-oki_ RSN4860_CHUETSU_65033NS.AT2 11 0.18
Sanjo Shinbori 6.8 23.18 0.21
28H2 Japan RSN4860_CHUETSU_65033EW.AT2 20 0.09
29H1 Joetsu RSN4846_CHUETSU_65009NS.AT2 4 0.28
Chuetsu-oki_
Yanagishima 6.8 31.43 0.39
29H2 Japan RSN4846_CHUETSU_65009EW.AT2 3 0.33
paddocks
30H1 COALINGA - RSN4150_PARK2004_46174-90.AT2 9 0.02
Parkfield-02_ CA PRIEST 6 22.02 0.05
30H2 RSN4150_PARK2004_46174360.AT2 10 0.03
VALLEY
31H1 Inglewood - RSN981_NORTHR_ING000.AT2 12 0.09
Northridge-01 6.69 42.2 0.12
31H2 Union Oil RSN981_NORTHR_ING090.AT2 10 0.10
32H1 San Bernardino - RSN930_BIGBEAR_SB2270.AT2 12 0.11
Big Bear-01 2nd & 6.46 33.79 0.13
32H2 RSN930_BIGBEAR_SB2360.AT2 10 0.10
Arrowhead
204

Table A 9 – Manuscript 3: Long- and short duration ground motions selected and respective
scale factors applied in IDA analysis
Pair Long-duration Short-duration
SF
# Earthquake Year Station File name Earthquake Year Station File name
Zack
Valparaiso El Chalfant RSN558_CHALFANT.A_A
1 1985 valparaiso_almendralL.txt 1986 Brothers 0.56
(Chile) Almendral Valley-02 -ZAK360.AT2
Ranch
San Ramon
Valparaiso Livermore-
2 1985 Llolleo llolleo1002271Lg.txt 1980 - Eastman LIVERMOR/B-KOD180.AT2 1.87
(Chile) 02
Kodak
Maule Irpinia,
3 2010 Angol angol1002271EWg.txt 1980 Brienza ITALY/A-BRZ000.AT2 2.89
(Chile) Italy-01
Managua,
Maule Managua,
4 2010 Constitucion constitucionLg.txt Nicaragua- 1972 MANAGUA_A-ESO090.AT2 1.93
(Chile) ESSO
01
Parkfield -
Maule
5 2010 Curico curicoNSg.txt Coalinga-01 1983 Stone COALINGA.H/H-SC3090.AT2 5.00
(Chile)
Corral 3E
Maule Santa Cachuma
6 2010 Hualane hualane1002271Tg.txt 1978 SBARB_CAD250.AT2 5.00
(Chile) Barbara Dam Toe
Parkfield -
Tohoku KR_FKSH19 Parkfield-02,
7 2011 Miyakoji 2004 Cholame PARK2004/C05090.AT2 2.47
(Japan) 1103111446EW2.AT2 CA
5W
Tohoku processed_YMTH06 Gilroy
8 2011 Takahata Coyote Lake 1979 COYOTELK/G04360.AT2 1.04
(Japan) 1103111446.EW2 Array #4
Tohoku processed_FKSH16 Friuli, Italy-
9 2011 Fukushima 1976 San Rocco FRIULI.B_B-SRO000.AT2 4.02
(Japan) 1103111446.EW2 02
Tohoku processed_MYGH08 L'Aquila,
10 2011 Iwanuma 2009 Celano L-AQUILA/TK003YLN.AT2 2.76
(Japan) 1103111446.EW2 Italy
Tohoku Imperial RSN167_IMPVALL.H_H
11 2011 Tsukidate MYG004Ewp 1979 Compuertas
(Japan) Valley-06 -CMP015.AT2
Tohoku
12 2011 Sakura CHB007NSp Tabas, Iran 1978 Dayhook RSN139_TABAS_DAY-L1.AT2
(Japan)
Pasadena -
Southern San
13 2010 Moquegua Moquegua_EW 1971 CIT RSN79_SFERN_PAS000.AT2
Peru Fernando
Athenaeum
Cholame -
Tohoku
14 2011 Haga TCGH16EWlp Parkfield 1966 Shandon RSN28_PARKF_C12050.AT2
(Japan)
Array #12
Southern Imperial Coachella RSN166_IMPVALL.H_H
15 2010 Arica Casa ACA_NS 1979
Peru Valley-06 Canal #4 -CC4045.AT2
Cedar
Maule
16 2010 Papudo PapudoL Lytle Creek 1970 Springs RSN42_LYTLECR_CSP126.AT2
(Chile)
Pumphouse
Tohoku San Golden
17 2011 Chiba CHB009_Ewp 1957 RSN23_SANFRAN_GGP010.AT2
(Japan) Francisco Gate Park
Tohoku Hollister RSN99_HOLLISTR_A-
18 2011 Hirata FKSH12_NS Hollister-03 1974
(Japan) City Hall HCH181.AT2
Maule Vina del San Lake
19 2010 VinaelsaltoEW 1971 RSN70_SFERN_L01021.AT2
(Chile) Mar Fernando Hughes #1
Chi-Chi Irpinia,
20 1999 CWB ALS chichi_NS 1980 Bisaccia RSN297_ITALY_B-BIS000.AT2
(Taiwan) Italy-02
Valparaiso Mammoth Convict RSN233_MAMMOTH.J_J
21 1985 Laligua CC06_200 1980
(Chile) Lakes-02 Creek -CVK090.AT2
Valparaiso San
22 1985 CC25_EW Gazli, USSR 1976 Karakyr RSN126_GAZLI_GAZ000.AT2
(Chile) Fernando
205

Manuscript 3: Copulas with correlations between starting data and reduced data used for

the bearing pressure model

(a) (b) (c)

(d) (e)
Figure A 1 - Copulas showing the parameters correlations used for footing #2 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
206

(a) (b) (c)

(d) (e)
Figure A 2 - Copulas showing the parameters correlations used for footing #3 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
207

(a) (b) (c)

(d) (e)
Figure A 3 - Copulas showing the parameters correlations used for footing #4 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
208

(a) (b) (c)

(d) (e)
Figure A 4 - Copulas showing the parameters correlations used for footing #5 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
209

(b) (b) (c)

(d) (e)
Figure A 5 - Copulas showing the parameters correlations used for footing #1 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
210

(b) (b) (c)

(d) (e)
Figure A 6 - Copulas showing the parameters correlations used for footing #2 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
211

(a) (b) (c)

(d) (e)
Figure A 7 - Copulas showing the parameters correlations used for footing #3 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
212

(b) (b) (c)

(d) (e)
Figure A 8 - Copulas showing the parameters correlations used for footing #4 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
213

(b) (b) (c)

(d) (e)
Figure A 9 - Copulas showing the parameters correlations used for footing #5 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
214

(a) (b) (c)

(d) (e)
Figure A 10 - Copulas showing the parameters correlations used for footing #1 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
215

(a) (b) (c)

(d) (e)
Figure A 11 - Copulas showing the parameters correlations used for footing #2 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
216

(a) (b) (c)

(d) (e)
Figure A 12 - Copulas showing the parameters correlations used for footing #3 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
217

(a) (b) (c)

(d) (e)
Figure A 13 - Copulas showing the parameters correlations used for footing #4 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
218

(a) (b) (c)

(d) (e)
Figure A 14 - Copulas showing the parameters correlations used for footing #5 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
219

(a) (b) (c)

(d) (e)
Figure A 15 - Copulas showing the parameters correlations used for footing #1 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
220

(a) (b) (c)

(d) (e)
Figure A 16 - Copulas showing the parameters correlations used for footing #2 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
221

(a) (b) (c)

(d) (e)
Figure A 17 - Copulas showing the parameters correlations used for footing #3 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
222

(a) (b) (c)

(d) (e)
Figure A 18 - Copulas showing the parameters correlations used for footing #4 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
223

(a) (b) (c)

(d) (e)
Figure A 19 - Copulas showing the parameters correlations used for footing #5 of the 3-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
224

(a) (b) (c)

(d) (e)
Figure A 20 - Copulas showing the parameters correlations used for footing #1 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
225

(a) (b) (c)

(d) (e)
Figure A 21 - Copulas showing the parameters correlations used for footing #2 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
226

(a) (b) (c)

(d) (e)
Figure A 22 - Copulas showing the parameters correlations used for footing #3 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
227

(a) (b) (c)

(d) (e)
Figure A 23 - Copulas showing the parameters correlations used for footing #4 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
228

(a) (b) (c)

(d) (e)
Figure A 24 - Copulas showing the parameters correlations used for footing #5 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
229

(a) (b) (c)

(d) (e)
Figure A 25 - Copulas showing the parameters correlations used for footing #6 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
230

(a) (b) (c)

(d) (e)
Figure A 26 - Copulas showing the parameters correlations used for footing #1 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
231

(a) (b) (c)

(d) (e)
Figure A 27 - Copulas showing the parameters correlations used for footing #2 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
232

(a) (b) (c)

(d) (e)
Figure A 28 - Copulas showing the parameters correlations used for footing #3 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
233

(a) (b) (c)

(d) (e)
Figure A 29 - Copulas showing the parameters correlations used for footing #4 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
234

(a) (b) (c)

(d) (e)
Figure A 30 - Copulas showing the parameters correlations used for footing #5 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
235

(a) (b) (c)

(d) (e)
Figure A 31 - Copulas showing the parameters correlations used for footing #6 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=15% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
236

(a) (b) (c)

(d) (e)
Figure A 32 - Copulas showing the parameters correlations used for footing #1 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
237

(a) (b) (c)

(d) (e)
Figure A 33 - Copulas showing the parameters correlations used for footing #2 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
238

(a) (b) (c)

(d) (e)
Figure A 34 - Copulas showing the parameters correlations used for footing #3 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
239

(a) (b) (c)

(d) (e)
Figure A 35 - Copulas showing the parameters correlations used for footing #4 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
240

(a) (b) (c)

(d) (e)
Figure A 36 - Copulas showing the parameters correlations used for footing #5 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
241

(a) (b) (c)

(d) (e)
Figure A 37 - Copulas showing the parameters correlations used for footing #6 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 10
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
242

(a) (b) (c)

(d) (e)
Figure A 38 - Copulas showing the parameters correlations used for footing #1 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
243

(a) (b) (c)

(d) (e)
Figure A 39 - Copulas showing the parameters correlations used for footing #2 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
244

(a) (b) (c)

(d) (e)
Figure A 40 - Copulas showing the parameters correlations used for footing #3 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
245

(a) (b) (c)

(d) (e)
Figure A 41 - Copulas showing the parameters correlations used for footing #4 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
246

(a) (b) (c)

(d) (e)
Figure A 42 - Copulas showing the parameters correlations used for footing #5 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
247

(a) (b) (c)

(d) (e)
Figure A 43 - Copulas showing the parameters correlations used for footing #6 of the 9-story
building considering 50,000 parameter sets, with variability associated to COV=30% and δh = 30
m. Points in red correspond to the reduced data set : (a) k1-k2, (b) k1-qSTC, and (c) k2-qSTC; (d) k1-
1/k2; (e) 1/k1-qSTC
248

Manuscript 3 - Code for implementation of HyperbolicQy material model


249
250
251
252
253

You might also like